
Citation: Yuxuan Zhang, Xinmiao Rong, Jimin Zhang. A diffusive predator-prey system with prey refuge and predator cannibalism[J]. Mathematical Biosciences and Engineering, 2019, 16(3): 1445-1470. doi: 10.3934/mbe.2019070
[1] | K. Maqbool, S. Shaheen, A. M. Siddiqui . Effect of nano-particles on MHD flow of tangent hyperbolic fluid in a ciliated tube: an application to fallopian tube. Mathematical Biosciences and Engineering, 2019, 16(4): 2927-2941. doi: 10.3934/mbe.2019144 |
[2] | Nattawan Chuchalerm, Wannika Sawangtong, Benchawan Wiwatanapataphee, Thanongchai Siriapisith . Study of Non-Newtonian blood flow - heat transfer characteristics in the human coronary system with an external magnetic field. Mathematical Biosciences and Engineering, 2022, 19(9): 9550-9570. doi: 10.3934/mbe.2022444 |
[3] | Chii-Dong Ho, Gwo-Geng Lin, Thiam Leng Chew, Li-Pang Lin . Conjugated heat transfer of power-law fluids in double-pass concentric circular heat exchangers with sinusoidal wall fluxes. Mathematical Biosciences and Engineering, 2021, 18(5): 5592-5613. doi: 10.3934/mbe.2021282 |
[4] | Chii-Dong Ho, Jr-Wei Tu, Hsuan Chang, Li-Pang Lin, Thiam Leng Chew . Optimizing thermal efficiencies of power-law fluids in double-pass concentric circular heat exchangers with sinusoidal wall fluxes. Mathematical Biosciences and Engineering, 2022, 19(9): 8648-8670. doi: 10.3934/mbe.2022401 |
[5] | Ziad Khan, Hari Mohan Srivastava, Pshtiwan Othman Mohammed, Muhammad Jawad, Rashid Jan, Kamsing Nonlaopon . Thermal boundary layer analysis of MHD nanofluids across a thin needle using non-linear thermal radiation. Mathematical Biosciences and Engineering, 2022, 19(12): 14116-14141. doi: 10.3934/mbe.2022658 |
[6] | Abdulaziz Alsenafi, M. Ferdows . Effects of thermal slip and chemical reaction on free convective nanofluid from a horizontal plate embedded in a porous media. Mathematical Biosciences and Engineering, 2021, 18(4): 4817-4833. doi: 10.3934/mbe.2021245 |
[7] | Aftab Ahmed, Javed I. Siddique . The effect of magnetic field on flow induced-deformation in absorbing porous tissues. Mathematical Biosciences and Engineering, 2019, 16(2): 603-618. doi: 10.3934/mbe.2019029 |
[8] | Yi Ren, Guolei Zhang, Longbin Yang, Yanwei Hu, Xiaojing Nie, Zhibin Jiang, Dawei Wang, Zhifan Wu . Study on seafloor hydrothermal systems circulation flow and heat transfer characteristics. Mathematical Biosciences and Engineering, 2022, 19(6): 6186-6203. doi: 10.3934/mbe.2022289 |
[9] | Niksa Praljak, Shawn D. Ryan, Andrew Resnick . Pulsatile flow through idealized renal tubules: Fluid-structure interaction and dynamic pathologies. Mathematical Biosciences and Engineering, 2020, 17(2): 1787-1807. doi: 10.3934/mbe.2020094 |
[10] | Jian Xu, Qian Shao, Ruo Chen, Rongrong Xuan, Haibing Mei, Yutao Wang . A dual-path neural network fusing dual-sequence magnetic resonance image features for detection of placenta accrete spectrum (PAS) disorder. Mathematical Biosciences and Engineering, 2022, 19(6): 5564-5575. doi: 10.3934/mbe.2022260 |
Since the pioneering works of Lorentz [1] and Oseen [2], the use of fundamental solutions has become a common and widely applied approach (referred to as the method of hydrodynamic singularities) for solving incompressible Stokes flows [3,4,5]. Important theoretical results in low-Reynolds number hydrodynamics have been obtained in this way, for instance, in quantifying the resistance of an arbitrarily shaped particle in a confined fluid [6], in constructing exact solutions for simple flows [7,8], in expressing the Generalized Faxén theorem [9] for generic immersed bodies. Furthermore, hydrodynamic singularities represent also one of the principal tools in numerical methods, such as Stokesian dynamics [10] or the Method of Fundamental Solutions [11].
Depending on the presence of a solid boundary at finite distance from the pole of the singularity, a distinction can be made between unbounded and bounded singularities [4]. In dealing with bounded singularities we consider, throughout this article, exclusively no-slip conditions at the boundaries.
In the unbounded case, all the hydrodynamic singularities can be constructed starting from the Green function for the Stokes flow, representing the lowest-order singularity, referred to as the Stokeslet [12], the Oseen tensor [3] or the Lorentzlet [13], see also [14] for a gradient-gauge approach, by applying to it a differential operator at the pole or at the source point. The relative simplicity in constructing hydrodynamic fields as a linear superposition of a collection of unbounded singularities has made the use of singular solutions extremely popular in the analytical description of velocity fields originated by the motion of solid bodies with different geometries in a Stokes fluid, thus simplifying considerably their representation with respect to those obtained by means of other approaches involving polar coordinates or multipole expansions [3]. Some well-known examples of solutions of hydrodynamic problems expressed in the singular representation refer to the motion of solid spheres [15], ellipsoids [7,8,9,16], tori [17] or slender bodies [12,18,19,20] in unbounded Stokes fluids. Moreover, the singular representation of the solutions of the Stokes flow has been used for characterizing the locomotion of microorganisms [12,21], and the rheological behavior of suspensions and complex fluids [22,23].
In the overwhelming majority of these works, the singularity functions are represented in a Cartesian reference system, since either the flow domain is unbounded, or, in the bounded case, the singularities lie on a flat manifold (mainly points and lines). In point of fact, a general theory of the Stokes singularities, should take into account any possible system configuration, that can, in principle, be constituted by curved boundaries (such as cylindrical channels, spheroidal capsules, wavy surfaces etc.), and immersed curved objects (helical flagellae, biconcave disk shaped cells etc.), for which it is convenient to associate singularities lying on curved manifolds due to their symmetries. Therefore, it may happen that the appropriate coordinate system for specific hydrodynamic problems is curvilinear. As well known, Navier-Stokes fields are invariant under coordinate transformations and, it easy to show, that hydrodynamic singularities are invariant also at the pole.
Tensor calculus [24] is the natural geometric framework for addressing invariance with respect to coordinate systems. In dealing with the singularity approach to Stokes flows, the singular fields depend at least on two points (and in principle, are multi-point functions), the source (at the pole of the singularity) and the field point (at the fluid element position). Consequently, a generalization of tensor calculus is required, represented by the bitensorial formalism [25,26,27], specifically developed for handling the Green functions in field theoretical developments within the theory of general relativity. The bitensor calculus, developed originally by Ruse [25], and further extended by Synge [26] and De Witt [27] for describing multi-point dependent fields in general relativity, is an extension of the tensor calculus that allows us to distinguish between the components of two-point dependent tensors (such as the Stokes singularities) and to make operations between them by means of the so called parallel propagator. A thorough analysis of bitensor calculus can be found in [28], while Appendix A succintly reviews the main concepts used in this article.
One goal of the present article is to develop a bitensorial formalism that ensures and preserves in a simple way invariant relations for the hydrodynamic singularity functions both at the source and the field points. In a broader perspective, the aim of this work is not only to transfer the bitensor formalism to the analysis of the hydrodynamic Green functions, which is a useful task in itself, as it makes the Stokesian formalism clear and unambiguous, but also to derive out of this formalism new hydrodynamic properties and operators. A significant example involves the generalization of the Faxén operator associated with an immersed body. In the hydrodynamic literature [3,4], the Faxén operator associated with an immersed body is defined as the integro-differential operator that, once applied to the ambient velocity field, provides the force acting on the body. Specifically, we derive the analytic expression for the generalized n-th order Faxén operator that, starting from the fundamental Green function (e.g., a Stokeslet), provides the disturbance field associated with an n-th order ambient flow (see Section 5 for details).
The article is organized as follows. Section 2 introduce the tensor algebra within the framework of the Stokes equations. In Section 3, we show how bitensor calculus eliminates the formal ambiguities (related to the meaning of the tensorial indices, and to the action of linear operators on tensorial singularities) occurring in the current formulation of Stokesian hydrodynamics [3,4] and it allows us to obtain a clear definition of singular solutions of Stokes flow (bounded and unbounded), specifying the associated homogeneous equations and boundary conditions.
Since the Stokes singularities can be viewed as generalized functions (or distributions), the generalized function theory [29] and its connections with the theory of moments [30] are applied to bitensorial quantities of hydrodynamic interest in Section 4. Specifically, we show that the linear operator providing the singularity system of a bounded flow is uniquely specified by the system of moments associated with the forces acting on the obstacle. Although the present definition of moments is altogether different from that proposed in [31], where, assuming no-slip boundary conditions, the moments are defined by surface integrals of the stress tensor, the two approaches yield the same final result as regard the expression of the disturbance field, showing that the no-slip boundary condition assumption is unnecessary. This represents the only intersection point between the present theory and the one developed by Ichiki [31] in the particular case of no-slip spheres in a Stokes flow. In Section 5, we introduce the n-th order Faxén operator, and we derive an n-th order Faxén theorem, by expressing a generic disturbance field as a series expansion involving the n-th order Faxén operators.
In Section 6, the operator yielding the disturbance field associated with a Stokeslet is considered showing that it is directly related to the reflection operator [32,33] of the geometry considered. This result is applied in Section 7 to the singularities near a plane wall. The characterization of the singularities bounded by a no-slip planar wall has been analyzed in the literature either as a reflection problem [1,34,35] or using a system of image singularities [36,37]. These two approaches are reviewed in [3]. We show in Section 7 that the present formalism highlights the equivalence between these two approches. In fact, the same differential operator furnishes directly either the Lorentz's mirror form of the solution, if applied at the field point, or the Blakes' singularity solution form, if applied at the source point of the Stokeslet. Moreover, since the position of the pole enters as a variable in the reflection operator, this formalism overcomes the original shortcomings in obtaining the higher order bounded singularities by differentiating the Green's function at the pole, due to the fact that, in the Blake's solutions, the distance of the pole from the plane enters as a parameter. In this way, we obtain unknown (Source Dipole and Stresslet) and known (Rotlet and Sourcelet) bounded singularities, the latter ones already derived in [37] by means of a more elaborate Fourier-Hankel transform.
If a Newtonian fluid, possessing viscosity μ, is subjected to a volume force field f(x), the controvariant components of the stress field σσ(x), the velocity v(x) and the scalar pressure field p(x) are solution, for vanishing Reynolds number, and under steady conditions, of the Stokes equations [38]
{−∇bσab(x)=μΔxva(x)−∇ap(x)=−fa(x)∇ava(x)=0x∈Vf | (2.1) |
a=1,2,3 where Vf is the fluid domain. Throughout this article, the Einstein summation convention is adopted. The operators ∇a and ∇a in Eq (2.1) represent the covariant and controvariant derivatives, respectively, related by the transformation ∇a=gab∇b, where gab=gab(x) is the metric tensor [39] and Δx=gab∇a∇b is the Laplacian operator at the point x. For a rank-2 tensor Ta b in mixed representation, its covariant derivative reads
∇cTa b=Ta b;c=∂Ta b∂xc+ΓamcTm b−ΓnbcTa n | (2.2) |
where Ta b;c is an alternative and more compact notation for the covariant derivative of Ta b, and Γamc are the Christoffel symbols
Γamc=12gal(∂glm∂xc+∂glc∂xm−∂gmc∂xl) | (2.3) |
Henceforth, we will use both the notations ∇cTa b and Ta b;c for the covariant derivatives.
The component of the associated stress tensor for a Newtonian incompressible fluid are therefore expressed by [38]
σab=pgab−μ(∇bva+∇avb)=pgab−μ(va;b+vb;a) | (2.4) |
As well known, the controvariant components of the generic tensorial field f(x)=(fa(x)) change from the coordinate system {xa} to a new system {˜xa} via a linear transformation defined by the matrix (∂˜xb∂xa)
˜fb(x)=fa(x)∂˜xb∂xa | (2.5) |
whereas the inverse matrix at the point x yields the transformation of the covariant components
˜fb(x)=fa(x)∂˜xa∂xb | (2.6) |
In this section we extend the tensorial notation to the case of the fundamental solutions of the Stokes flow, with the aim of obtaining a clear definition of its singular solutions From the theory of distributions, we can write the fields entering Eq (2.1) equipped with homogeneous Dirichlet boundary condition at ∂Vf as volume potentials [40], with a kernel Ga α(x,ξξ) for the velocity field
va(x)=∫Ga α(x,ξξ)fα(ξξ)8πμ√g(ξξ)d3ξ | (3.1) |
and a kernel Pα(x,ξξ) for the pressure field
p(x)=∫Pα(x,ξξ)fα(ξξ)8π√g(ξξ)d3ξ | (3.2) |
where fα(ξξ) are the controvariant components of the force field at a source point ξξ, g(ξξ)=det(gab(ξξ)) and d3ξ=dξ1dξ2dξ3. Observe that the coordinate representation of the source point ξξ could in principle be different from that of the field point x. This fact is notationally highlighted throughout the article, by using greek letters instead of latin ones for any index α=1,2,3 referred to the entries of tensorial entities evaluated at the source point. Therefore, the transformations for the controvariant and covariant components of f(ξξ) at the source point read
fβ′(ξξ′)=fα(ξξ)∂ξβ′∂ξα,fβ′(ξξ)=fα(ξξ)∂ξα∂ξβ′ | (3.3) |
where ξβ′ are the components of ξξ′. This notation, with primed indices to indicate the transformed coordinated, will be used throughout the article.
The kernels Ga α(x,ξξ) and Pα(x,ξξ) are two-point dependent distributions, with tensorial character both at x and ξξ, thus corresponding to bitensorial quantities [25,26,27,28]. This is a common feature of any fundamental solutions (or Green functions) in mathematical physics. Further details on the theory of bitensors are succintly reviewed in Appendix A. Specifically, the kernel entries Ga α(x,ξξ) are the components of a bitensor with vectorial character both at the source and the field point, and consequently their transformation in new coordinate systems both at the source and the field points takes the form
Gb′ β′(x′,ξξ′)=Ga α(x,ξξ)∂xb′∂xa∂ξα∂ξβ′ | (3.4) |
whereas the transformation rule for the pressure bitensor, with scalar character at the field point x and vectorial at the source point ξξ, is given by
Pβ′(x,ξξ′)=Pα(x,ξξ)∂ξα∂ξβ′ | (3.5) |
Finally, using the invariance properties of the Dirac delta function [28] and the parallel transport of tensorial quantities, it is possible to express the force field entering Eq (2.1) as
fa(x)=∫ga α(x,ξξ)fα(ξξ)δ(x,ξξ)√g(ξξ)d3ξ;δ(x,ξξ)=δ(x−ξξ)√g(ξξ) | (3.6) |
where ga α(x,ξξ) is the parallel propagator bitensor, which propagates in a parallel way a vector along the unique geodesics connecting x to ξξ. In a distributional meaning, it follows that gaα(x,ξξ)δ(x,ξξ)=δa αδ(x,ξξ) being x and ξξ coincident.
By substituting Eqs (3.1), (3.2) and (3.6) in Eq (2.1), we obtain the bitensorial Green function equations of the Stokes flow, yielding the velocity and pressure at the field point x due to an impulsive force acting at the source point ξξ
{−∇bΣabα(x,ξξ)=ΔxGa α(x,ξξ)−∇aPα(x,ξξ)=−8πδaαδ(x,ξξ)∇aGa α(x,ξξ)=0Ga α(x,ξξ)|x∈∂Vf=0 | (3.7) |
From Eq (2.4), the stress field Σabα(x,ξξ) associated with the Green function is defined by
Σabα(x,ξξ)=Pα(x,ξξ)gab(x)−(Ga;b α(x,ξξ)+Gb;a α(x,ξξ)) | (3.8) |
In the case the source point is kept fixed, bitensors become simple tensors depending only on the field point. Therefore, by choosing the force field f(ξξ)=f0δ(ξξ−ξξ0), from Eqs (3.1) and (3.2), we obtain the velocity/pressure fields due to an impulsive force with intensity f0 placed at a singular point ξξ0
va(x)=fα08πμGa α(x,ξξ0) | (3.9) |
p(x)=fα08πPα(x,ξξ0) | (3.10) |
for which the stress tensor σab(x) takes the form
σab(x)=fα08πΣab α(x,ξξ0) | (3.11) |
The reciprocity relation [3,4] for the Green function in bitensorial notation becomes
Ga α(x,ξξ)=G aα(ξξ,x) | (3.12) |
By exchanging x↔ξξ, and thus a↔α, and enforcing the reciprocity relation (3.12), it follows that Ga α(x,ξξ) is also the solution of the system
{−∇βΣaαβ(ξξ,x)=ΔξGa α(x,ξξ)−∇αPa(ξξ,x)=−8πδaαδ(x,ξξ)∇αGa α(x,ξξ)=0Ga α(x,ξξ)|ξξ∈∂Vf=0 | (3.13) |
where Δξ=gαβ(ξξ)∇α∇β is the Laplacian at point ξξ. In this case, the associated stress field becomes
Σaαβ(ξξ,x)=Pa(ξξ,x)gαβ(ξξ)−(Ga α;β(x,ξξ)+Ga β;α(x,ξξ)) | (3.14) |
Since the Green function vanishes at ξξ∈∂Vf for any x, Pα(x,ξξ) must be constant for ξξ∈∂Vf due to Eq (3.7), and therefore can be set equal to zero. Furthermore, the pressure scalar-vector Pα(x,ξξ) is a potential scalar field at x possessing the following properties
ΔxPα(x,ξξ)=8π∇αδ(x,ξξ)∇αPα(x,ξξ)=8πδ(x,ξξ)Pα(x,ξξ)|ξξ∈∂Vf=0 | (3.15) |
The first relation stems from Eq (3.7), by taking the divergence with respect to x, while the second relation follows by taking the divergence with respect to ξξ, enforcing the second relation in Eq (3.13). In a similar way, Pa(ξξ,x) fulfills the relations
ΔξPa(ξξ,x)=8π∇aδ(x,ξξ)∇aPa(ξξ,x)=8πδ(x,ξξ)Pa(ξξ,x)|x∈∂Vf=0 | (3.16) |
Observe that Eq (3.15) for Pα(x,ξξ), and likewise Eq (3.16) for Pa(ξξ,x) do not constitute a boundary value problem for the pressure variable, as the boundary condition is assigned for a variable (ξξ in Eq (3.15)) different from that involved in the differential equation (x in Eq (3.15)), thus representing a collection of properties fulfilled by the pressure field.
To obtain the higher order singularities, the Green function should be differentiated at the pole ξξ maintaining homogeneous Dirichlet conditions at the field point. The first derivative at the pole yields the Stokesian dipole, the second derivative the Stokesian quadrupole and so on.
The Stokesian dipole Ga α;β(x,ξξ) can also be expressed as superposition of two other singular solutions of the Stokes equations: a symmetric and an antisymmetric tensor field at the source points
Ga α;β(x,ξξ)=Ea αβ(x,ξξ)+εγαβΩaγ(x,ξξ) | (3.17) |
where
Ea αβ(x,ξξ)=12(Ga α;β(x,ξξ)+Ga β;α(x,ξξ)) | (3.18) |
is the field due to a singular strain of the fluid at the source point, and
Ωaγ(x,ξξ)=εγϵη2Ga ϵ;η(x,ξξ) | (3.19) |
where εαβγ is the Levi-Civita symbol (in the italian mathematical literature also called the Ricci tensor [38]), is the field due to a singular rotation of the fluid at the source point.
The symmetric strain component is the solution of the Stokes system of equations
{ΔxEa αβ(x,ξξ)−12∇a(Pα;β(x,ξξ)+Pβ;α(x,ξξ))=−4π(δaα∇β+δaβ∇α)δ(x,ξξ)∇aEa αβ(x,ξξ)=0Ea αβ(x,ξξ)|x∈∂Vf=0 | (3.20) |
which can be also computed directly from Eqs (3.13), (3.14) by exchanging source and field points in the pressure and stress fields related to the solution of the Green function
Ea α;β(x,ξξ)=gαβ(ξξ)2Pa(ξξ,x)−12Σaαβ(ξξ,x) | (3.21) |
The antisymmetric part of the Stokes dipole corresponds to the solution of the Stokes system
{ΔxΩaγ(x,ξξ)−12εγϵη∇aPϵ;η(x,ξξ)=−4πδaϵεγϵη∇ηδ(x,ξξ)∇aΩaγ(x,ξξ)=0Ωaγ(x,ξξ)|x∈∂Vf=0 | (3.22) |
A further differentiation at the pole defines the Stokes quadrupole. Specifically, by applying the Laplacian operator Δξ/2 to the Green function, we obtain the so called Source Dipole
{ΔxDa α(x,ξξ)−12∇aΔξPα(x,ξξ)=−4πδaαΔξδ(x,ξξ)∇aDa α(x,ξξ)=0Da α(x,ξξ)|x∈∂Vf=0 | (3.23) |
Also the solution of this system can be obtained by exchanging source and field points in the gradient of the pressure field associated with the Green function. In point of fact, from the first relation in Eq (3.13), we have
Da α(x,ξξ)=−ΔξGa α(x,ξξ)2=−∇αPa(ξξ,x)2+4πδaαδ(x,ξξ) | (3.24) |
In this paragraph, unbounded singularities are briefly analyzed. Due to translational invariance, the singularities in R3, depend solely on the vector x−ξξ. Henceforth, the unbounded singular functions will be indicated by sans-serif capital letters. The Green function Sa α(x−ξξ), usually referred to as the Stokeslet, is the solution of the Stokes problem
{−∇bΣabα(x−ξξ)=ΔxSa α(x−ξξ)−∇aPα(x−ξξ)=−8πδaαδ(x,ξξ)∇aSa α(x−ξξ)=0Sa α(x−ξξ)||x−ξξ|→∞=0 | (3.25) |
Since the Laplacian is invariant under translation (and, more generally, under Euclidean transformations [41]), we have for a generic function f(x−ξξ), Δx−ξf(x−ξξ)=Δxf(x−ξξ)=Δξf(x−ξξ). Therefore, it is possible to express the pressure in Eq (3.15) as the solution of the harmonic problem
{ΔξPα(x−ξξ)=8π∇αδ(x,ξξ)Pα(x−ξξ)||x−ξξ|→∞=0 | (3.26) |
thus
Pα(x−ξξ)=2∇α1|x−ξξ|=2(x−ξξ)αr3 | (3.27) |
while that the associated velocity and stress-tensor fields are given by [4,14]
Sa α(x−ξξ)=(δaαΔξ−∇α∇a)|x−ξξ|=δaα|x−ξξ|+(x−ξξ)a(x−ξξ)α|x−ξξ|3 | (3.28) |
Σabα(x−ξξ)=6(x−ξξ)a(x−ξξ)b(x−ξξ)αr5 | (3.29) |
As Pα;β(x−ξξ)=Pβ;α(x−ξξ), the symmetric part of the Stokes dipole corresponds to the solution of the problem
{ΔxEa αβ(x−ξξ)−∇aPα;β(x−ξξ)=−4π(δaα∇β+δaβ∇α)δ(x,ξξ)∇aEa αβ(x−ξξ)=0Ea αβ(x−ξξ)||x−ξξ|→∞=0 | (3.30) |
and due to Eq (3.21) it takes the expression
Ea αβ(x−ξξ)=gαβ(ξξ)2Pa(ξξ−x)−12Σaαβ(ξξ−x) | (3.31) |
This field can be viewed as the superposition of two terms: the contribution Ma(x−ξξ)=−Pa(ξξ−x)/2, which is the solution of a Stokes problem everywhere but at the pole
{ΔxMa(x−ξξ)=−4π∇aδ(x,ξξ)∇aMa(x−ξξ)=−4πδ(x,ξξ)Ma(x−ξξ)||x−ξξ|→∞=0 | (3.32) |
Strictly speaking, the field Ma(x−ξξ), usually called the Sourcelet [4,37], is not a Stokesian singular solution, since its divergence does not vanish at the pole and, thus, it does not satisfy the overall mass balance over the fluid. It can be physically interpreted as the velocity field stemming from a pointwise fluid source (or sink, if the sign is reversed) at the pole. Its bounded counterpart can be defined solely for external problems, so that it could match the regularity condition and the overall mass balance at infinity. However, it cannot be generally neither obtained from the Green function (as the Green function is divergence-free), nor it is related to the Green function pressure field, as in the unbounded case.
Similarly, also the second term is not a singular Stokesian solution. In fact, the field Taαβ(x−ξξ)=Σaαβ(ξξ−x)/2, called the Stresslet, is the solution of the problem
{ΔxTaαβ(x−ξξ)−∇aPα;β(x−ξξ)=−4π(gαβ(ξξ)∇a+δaα∇β+δaβ∇α)δ(x,ξξ)∇aTaαβ(x−ξξ)=−4πgαβ(ξξ)δ(x,ξξ)Taαβ(x−ξξ)||x−ξξ|→∞=0 | (3.33) |
possessing non vanishing divergence. Therefore, the symmetric Strainlet Eq (3.31) can be expressed as
Ea αβ(x−ξξ)=−gαβ(ξξ)Ma(x−ξξ)+Taαβ(x−ξξ) | (3.34) |
Next consider the antisymmetric term defined by Eq (3.19). In unbounded flows Ωaγ(x−ξξ) is referred to as the Rotlet. Since εγϵηPϵ;η(x−ξξ)=0, the Rotlet is a constant pressure solution of the Stokes system
{ΔxΩaγ(x−ξξ)=−4πδaϵεγϵη∇ηδ(x,ξξ)∇aΩaγ(x−ξξ)=0Ωaγ(x−ξξ)||x−ξξ|→∞=0 | (3.35) |
the analytic expression of which is
Ωaγ(x−ξξ)=−δaϵεγϵη∇η1|x−ξξ| | (3.36) |
Another low order irrotational singularity of the Stokes problem is the solution of Eq (3.23) in unbounded domain, namely
{ΔxDa α(x−ξξ)=−4π(δaαΔξ−∇a∇α)δ(x,ξξ)∇aDa α(x−ξξ)=0Da α(x−ξξ)||x−ξξ|→∞=0 | (3.37) |
This solution, referred to as the Source Doublet, can be obtained from Eq (3.25) and from the definition of Ma(x−ξξ)
Da α(x−ξξ)=−ΔξSa α(x−ξξ)2=∇αMa(x−ξξ)+4πδaαδ(x,ξξ) | (3.38) |
In the previous section we have discussed how all the singularities of bounded flows can be obtained by differentiating the Stokeslet at its pole. In this Section, we develop a method to obtain the singular representation of a Stokes flow in a given domain Vf containing solid boundaries by means of a linear operator applied to the Stokeslet in the external domain Vext≡R3/Vf, and yielding the disturbance field in Vf. More precisely, consider a given solution u(x) of the Stokes equation in Vf, attaining arbitrary values at the boundaries ∂Vf (at which, the Stokes problem dictates no-slip boundary conditions). The velocity field u(x) is referred to as the ambient flow. In order to match the no-slip boundary condition, a disturbance flow w(x) should be added so that v(x)=w(x)+u(x) is the Stokes solution within Vf satisfying the no-slip conditions on ∂Vf. Thus, the disturbance flow is a solution of the equations
{μΔxwa(x)−∇aq(x)=0x∈Vf∇awa(x)=0wa(x)=−ua(x)x∈∂Vf | (4.1) |
where q(x) is the associated pressure field. It is convenient to extend this problem over the whole physical space R3 in order to obtain its singular representation. To this purpose, we can formulate the problem defined by Eq (4.1) in the form of the non-homogeneous unbounded Stokes equations in R3 as
{μΔxwa(x)−∇aq(x)=−fa(x)x∈R3∇awa(x)=0 | (4.2) |
with the condition that fa(x) are distributions defined on a compact support in Vext, and satisfying the integral equation
∫VextSa α(x,ξξ)fα(ξξ)8πμ√g(ξξ)d3ξ=−ua(x)x∈∂Vf | (4.3) |
Let us introduce the n-th order tensorial moments of the function f(x), extending the scalar moment theory [30], as
Mα ααn(ξξ)=∫Vextgα a(ξξ,x)g a1α1(ξξ,x)...g anαn(ξξ,x)fa(x)(x−ξξ)a1...(x−ξξ)an√g(x)d3x,ξξ∈Vext | (4.4) |
or, using the scalar-product notation on the external domain
Mα ααn(ξξ)=⟨gα a(ξξ,x)fa(x),g anααn(ξξ,x)(x−ξξ)an⟩ | (4.5) |
where ⟨⋅,⋅⟩ indicates the scalar product in Vext≡R3/Vf, an=(a1,…,an) is a multi-index, g anααn(ξξ,x)=g a1α1(ξξ,x)...g anαn(ξξ,x) and (x−ξξ)an=(x−ξξ)a1...(x−ξξ)an. It is shown in Appendix B that the moments Ma an(ξξ) can be reduced to the Ichiki's surface integrals [31] in the case that the no-slip boundary conditions are assumed, thus
Mα ααn(ξξ)=−∫∂Vfgα a(ξξ,x)g anααn(ξξ,x)(x−ξξ)anσab(x)nb(x)dS(x) | (4.6) |
where σσ(x) is the stress tensor related to the total velocity field v(x), and nb(x) the covariant components of the outwardly oriented normal unit vector at points x of ∂Vf as shown in Figure 1. Therefore, given a reference point ξξ, all the moments on the volume Vext are uniquely determined by the stress field at the surface, since Eq (4.6) does not depend on the chosen function f(x).
Consider the tensorial Taylor expansion [42] of the components of the vectorial test function ϕϕ(x) around a given point ξξ∈Vext
ϕa(x)=∞∑n=0(−1)nga α(x,ξξ)∇ααnϕα(ξξ)n!(ξξ−x)ααn | (4.7) |
where ∇ααn=∇α1⋯∇αn. Owing to the bitensorial notation, there is no ambiguity in the definition of ∇ααn as greek indices refer to the source point.
Applying the test function to the momentum balance equation entering Eq (4.2) we have
⟨fa,ϕa⟩=∞∑n=0⟨fa,(−1)ng αa(x,ξξ)∇ααnϕα(ξξ)n!(ξξ−x)ααn⟩=∞∑n=0∇ααnϕα(ξξ)n!Mα ααn(ξξ) | (4.8) |
where we have made use of the relations g αa(x,ξξ)=gα a(ξξ,x) and (x−ξξ)ααn=(−1)n(ξξ−x)ααn=(x−ξξ)ang ααnan(ξξ,x) see Appendix A.
Since the derivatives of the test functions can be formulated in scalar-product notation as
∇ααnϕα(ξξ)=(−1)n⟨∇ααnga α(x,ξξ)δ(x,ξξ),ϕa(x)⟩ | (4.9) |
substituting Eq (4.9) into Eq (4.8), the function f(x) can be finally expressed as
fa(x)=∞∑n=0(−1)nMα ααn(ξξ)n!∇ααnga α(x,ξξ)δ(x,ξξ) | (4.10) |
Although the moments depend on the reference points ξξ, the summation in Eq (4.10) does not depend on ξξ. Therefore, Eq (4.10) can be generalized by considering ξξ as a point of an arbitrary k-dimensional (k≤3) set of points Ω, averaging Eq (4.10) over Ω,
fa(x)=1meas(Ω)∫ΩdΩ(ξξ)∞∑n=0(−1)nMα ααn(ξξ)n!∇ααnga α(x,ξξ)δ(x,ξξ) | (4.11) |
where dΩ(ξξ) is the measure element and
meas(Ω)=∫ΩdΩ(ξξ) |
is the Lebesgue measure of Ω. Depending on the symmetries of the flow geometry, the set Ω can be chosen in some particular cases as to reduce the infinite summation entering Eq (4.11) to a finite number of terms.
From the structure of Eq (4.10) we can introduce a differential operator
D∗α=∞∑n=0(−1)nMα ααn(ξξ)n!∇ααn | (4.12) |
that in Eq (4.10) acts on the Dirac delta function. In a similar way, if Eq (4.10) is generalized by Eq (4.11), the operator D∗α attains an integro-differential representation
D∗α=1meas(Ω)∫ΩdΩ(ξξ)∞∑n=0(−1)nMα ααn(ξξ)n!∇ααn | (4.13) |
so that fa=D∗αga α(x,ξξ)δ(x,ξξ). Its adjoint D, ⟨D∗af,g⟩=⟨f,Dg⟩, is expressed by
Dα=1meas(Ω)∫ΩdΩ(ξξ)∞∑n=0Mα ααn(ξξ)n!∇ααn | (4.14) |
Therefore, the problem defined by Eq (4.2) can be reformulated as
{μΔxwa(x)−∇aq(x)=−D∗αga α(x,ξξ)δ(x,ξξ)x∈R3,ξξ∈Vext∇awa(x)=0 | (4.15) |
and the singular representation of the velocity field w(x) follows from Eqs (3.1) and (3.2), namely
wa(x)=⟨D∗αδ(ξξ′,ξξ),Sa α(x−ξξ′)8πμ⟩=DαSa α(x−ξξ)8πμ | (4.16) |
and
p(x)=⟨D∗αδ(ξξ′,ξξ),P α(x−ξξ′)8π⟩=DαP α(x−ξξ)8π | (4.17) |
where the scalar products in Eqs (4.16), (4.17) correspond to an integration over ξξ′. Thus the operator D defined by the (4.14) provides the singular expansion, of the flow at the source point ξξ.
The existence of an integro-differential operator yielding the disturbance field once applied to the Stokeslet is hypothesized in Stokesian hydrodynamics for solid bodies [9,43] and, more generally, in developing the singularity method [4]. The procedure outlined above, based on the generalized function theory, provides an explicit expression for this operator in the form of a series expansion the coefficients of which are the moments. The main advantages of this explicit representation are: (i) for a specific flow problem the terms in the series expansion of the operator can be obtained numerically with arbitrary precision, (ii) it is possible to manipulate its formal structure in order to obtain new relations as will be shown in the next Sections.
In this section we focus on the Stokes flow around an object with no-slip boundary condition, thus considering the external volume Vext≡VB, where VB is the domain occupied by the object, bounded by the closed surface ∂VB. In this case Vf=R3/VB. For this class of hydrodynamic problems, the Faxén operator, and its generalizations, play an important role. In fact, as shown by Kim [9], the linear operator that, applied at the pole of the Stokeslet gives the velocity field due to a particle translating into the fluid, coincides with the operator that, applied to an ambient flow, returns the force onto the particle immersed in the flow. This operator takes its name from Faxén (see [23]), who derived its expression for the first time providing the force on a sphere from the values of the field and its Laplacian at the center of a spherical object. A corresponding theorem holds also for rotations and strains of the body.
Faxén operators for objects with geometry different from the sphere can be expressed in infinite series involving the derivatives of any order of the field evaluated at points within the domain corresponding to the object [44]. Exploiting the symmetry of particular objects, such as spheroids [7,8], ellipsoids [3], tori [17], etc., it is possible to express the same operator by means of a finite system of derivatives of the field evaluated on a manifold Ω⊂VB, such as the focal axis, the focal ellipse, the symmetry circle, etc..
Making use of the concepts and the formalism developed in Sections 3 and 4, and considering the Stokes flow around an object, below we derive the properties and the analytical representation of a generalization of the Faxén operator, referred to as the n-th order Faxén operator. To this aim, let us consider the disturbance field w(n)(x;ξξ′) of a purely n-th order unbounded ambient flow centered at the point ξξ′,
ua(n)(x;ξξ′)=Aa an(x−ξξ′)an,Aa an∇a[(x−ξξ′)an]=0 | (5.1) |
with pressure
p(n)(x;ξξ′)=μAa anˆpana(x;ξξ′),ˆpana(x;ξξ′)=n(x−ξξ′)an−2(x−ξξ′)aganan−1(x) | (5.2) |
and stress field
πbc(n)(x;ξξ′)=μAa anˆπbcana(x;ξξ′),ˆπbcana(x;ξξ′)=n((x−ξξ′)an−2(x−ξξ′)aganan−1(x)gbc(x)−(δbagcan(x)+δcagban(x))(x−ξξ′)an−1) | (5.3) |
i.e., the solution (w(n)(x),ττ(n)(x)) of the following Stokes problem
{μΔxw(n)(x;ξξ′)−∇q(n)(x;ξξ′)=−∇⋅ττ(n)(x;ξξ′)=0∇⋅w(n)(x;ξξ′)=0wa(n)(x;ξξ′)=−Aa an(x−ξξ′)anx∈∂Vf | (5.4) |
where the subscript "(n)" is not a tensorial index, but simply indicates that an n-order ambient flow is considered. As in Section 4, let us introduce the m-th order moments, associated with the stress tensor σσ(n)(x;ξξ′)=ττ(n)(x;ξξ′)+ππ(n)(x;ξξ′)
(n)Mα ααm(ξξ;ξξ′)=−∫∂Vfgα a(ξξ,x)g amααm(ξξ,x)(x−ξξ)amσab(n)(x;ξξ′)nb(x)dS(x) | (5.5) |
From the definition and the linearity of the problem, the moments in Eq (5.5) can be rewritten as,
(n)Mα ααm(ξξ;ξξ′)=8πμA ββnβmα β ααm ββn(ξξ;ξξ′) | (5.6) |
where the latter equation can be viewed as the definition of the geometric moments mα β ααm ββn(ξξ;ξξ′). In fact, enforcing linearity, i.e., σac(x;ξξ′)=A bnbˆσacbbn(x;ξξ′), one obtains for the geometric moments the following expression
mα β ααm ββn(ξξ;ξξ′)=−18π∫∂Vfgα a(ξξ,x)gβ b(ξξ,x)g amααm(ξξ,x)g bnββn(ξξ,x)(x−ξξ)amˆσacbbn(x;ξξ′)nc(x)dS(x) | (5.7) |
Geometric moments possess the following symmetry
mα β ααm ββn(ξξ;ξξ′)=mβ α ββn ααm(ξξ′;ξξ) | (5.8) |
that can be proved by applying the Lorentz reciprocal theorem
∫∂Vfw(m)(x,ξξ)⋅σσ(n)(x,ξξ′)⋅n(x)dS=∫∂Vfw(n)(x,ξξ)⋅σσ(m)(x,ξξ′)⋅n(x)dS | (5.9) |
Componentwise Eq (5.9) reads
A amaA bnb∫∂Vf(x−ξξ)amˆσacbbn(x;ξξ′)nb(x)dS(x)=A bnbA ama∫∂Vf(x−ξξ′)bnˆσbcaam(x;ξξ)nc(x)dS(x) | (5.10) |
from which Eq (5.8) follows.
Let us define the following integro-differential operator
Fαβ ββn=1meas(Ω)∫ΩdΩ(ξξ)∞∑m=0mα β ααm ββn(ξξ;ξξ′)m!∇ααm | (5.11) |
the adjoint of which is given by
F∗αβ ββn=1meas(Ω)∫ΩdΩ(ξξ)∞∑m=0(−1)mmα β ααm ββn(ξξ;ξξ′)m!∇ααm | (5.12) |
and introduce also its contracted form
Fα(n)=8πμA ββnβFαβ ββn | (5.13) |
From the analysis developed in Section 4, the Stokes problem (5.2) can be extended in R3 in the singular form as
{μΔxwa(n)(x;ξξ′)−∇qa(n)(x;ξξ′)=−F∗α(n)ga α(x,ξξ)δ(x,ξξ)∇⋅w(n)(x;ξξ′)=0 | (5.14) |
and from Eqs. (4.13)–(4.16), the velocity field solution of Eq (5.14) is given by
wa(n)(x;ξξ′)=Fα(n)Sa α(x−ξξ)8πμ | (5.15) |
Since it is possible to express the ambient velocity field u(n)(ξξ), ξξ∈VB, within the domain of the body in the surface integral form [4,40]
uα(ξξ)=−∫∂VBSα a(ξξ−x)8πμσab(x)nb(x)dS(x) | (5.16) |
by applying the operator F(n)α to the above ambient flow we obtain
F(n)αuα(ξξ)=−A ana∫∂Vf(x−ξξ)anσab(x)nb(x)dS(x) | (5.17) |
which, due to Eq (4.6), implies
Mα ααn(ξξ)=8πμFβα ααnuβ(ξξ) | (5.18) |
Therefore, the integro-differential operator defined by Eq (5.11) returns either the n-th order disturbance flow, once applied to the Stokeslet according to Eq (5.15), or the n-th order moment (4.6) once applied to the ambient flow according to Eq (5.18). In virtue of these properties, we can refer to Fβα ααn as the generalized n-th order Faxén operator.
We can now express the disturbance velocity field entering Eq (4.16), in the presence of generic Dirichlet boundary condition on the surface of the body, as
wa(x)=1meas(Ω)∫ΩdΩ(ξξ)∞∑n=0Fβα ααnuβ(ξξ)n!∇ααnSa α(x−ξξ) | (5.19) |
Expliciting the Faxén operator, and making use of the symmetries of the geometric moments entering Eq (5.8), we obtain
wa(x)=1meas(Ω)2∫ΩdΩ(ξξ)∫ΩdΩ(ξξ′)∞∑n=0∞∑m=0mβ α ββm ααn(ξξ′;ξξ)∇ββmuβ(ξξ′)m!n!∇ααnSa α(x−ξξ)=1meas(Ω)2∫ΩdΩ(ξξ)∫ΩdΩ(ξξ′)∞∑n=0∞∑m=0mα β ααn ββm(ξξ;ξξ′)∇ββmuβ(ξξ′)m!n!∇ααnSa α(x−ξξ) | (5.20) |
and since
Fαβ ββm=1meas(Ω)∫ΩdΩ(ξξ)∞∑n=0mα β ααn ββm(ξξ;ξξ′)∇ααnn! | (5.21) |
the disturbance velocity field can be expressed as a series expansion involving the generalized Faxén operators
wa(x)=1meas(Ω)∫ΩdΩ(ξξ′)∞∑m=0∇ββmuβ(ξξ′)m!Fαβ ββmSa α(x−ξξ) | (5.22) |
Equation (5.22) can be rewritten in a more compact form as
wa(x)=18πμ∞∑m=0Fα(m)Sa α(x−ξξ) | (5.23) |
where Fα(m)=8πμA ββmβFαβ ββm, as in the definition Eq (5.13), and the expansion coefficients A ββmβ are given by
A ββmβ=1meas(Ω)∫Ω∇ββmuβ(ξξ)m!dΩ(ξξ) | (5.24) |
Equation (5.23) provides a compact and elegant way to decompose the disturbance field due to a body immersed in an ambient flow in elementary motions associated with the different n-order velocity fields defined at the boundary of the object.
In hydrodynamics problems involving bounded flows and confined geometries, the Green function Ga α′(x,ζζ), solution of the equations
{−∇bΣabα′(x,ζζ)=ΔxGa α′(x,ζζ)−∇aPα′(x,ζζ)=−8πga α′(x,ζζ)δ(x,ζζ)∇aGa α′(x,ζζ)=0;x,ζζ∈VfGa α′(x,ζζ)=0;x∈∂Vf | (6.1) |
(referred for short to as the bounded Green function) plays a central role as it provides the volume potential in the fluid domain Vf, starting from which any flow with no-slip boundary conditions at ∂Vf, can be constructed.
Bounded Green function are available in the literature for a handful of simple geometries, as reviewed in [4]. In special cases, such as for the Green function of a fluid bounded by a plane [36] or outside a sphere [45,46] (see also [3]), a representation of the bounded Green function in terms of unbounded singularities placed outside the fluid domain is available. This representation is referred to as the image system [3], which is particularly handy for analytical and numerical calculations whenever the set of singularities is either finite or localized on simple manifolds. The latter property characterizes flows with suitable and simple symmetries while, for generic bounded flows, an image system of singularities is not available.
Based on the theory developed in Section 4, this Section addresses the properties of the operator providing the image system for a generic bounded Green function. To this aim, let us to consider as the ambient field the unbounded flow due to a Stokeslet centered at the point ζζ∈Vf and let us use the primed indices, say α′,β′,..., for referring to this point
ua(x)=fα′08πμSa α′(x−ζζ) | (6.2) |
As a consequence, the boundary condition for the disturbance field is given by
wa(x)=−ua(x)=−fα′08πμSa α′(x−ζζ),x∈∂Vf | (6.3) |
Owing to linearity, let us define the field Wa α′(x,ζζ), depending on ζζ, but regular at this point, such that
wa(x)=fα′08πμWa α′(x,ζζ) | (6.4) |
The theory developed in Section 4 can be applied, and enforcing Eq (4.16) the field Wa α′ is given by
Wa α′(x,ζζ)=Dα α′Sa α(x−ξξ) | (6.5) |
where
Dα α′=1meas(Ω)∫ΩdΩ(ξξ)∞∑n=0Mαα′ααn(ξξ,ζζ)n!∇ααn,Mαα′ααn(ξξ,ζζ)=−∫∂Vf(x−ξξ)ααngα a(ξξ,x)Σabα′(x,ζζ)8πnb(x)dS(x) | (6.6) |
Therefore, the Green function solution of Eq (6.1) can be expressed as the sum of two contributions: a singular part, due to the Stokeslet centered in the point ζζ, and a regular part due to the integro-differential operator Dα α′ acting on the poles of the Stokeslet outside the domain of the fluid
Ga α′(x,ζζ)=Sa α′(x−ζζ)+Dα α′Sa α(x−ξξ) | (6.7) |
Owing to the properties of the Green functions, the same result can be obtained by applying the operator Dα α′ at the field point. In point of fact, making use of the reciprocal identities for the Green functions, Ga α′(x,ζζ)=G aα′(ζζ,x) and Sa α′(x,ζζ)=S aα′(ζζ,x), it follows that
Ga α′(x,ζζ)=G aα′(ζζ,x)=S aα′(ζζ−x)+D aαS αα′(ζζ−ξξ)=Sa α′(x−ζζ)+D aαSα α′(ξξ−ζζ) | (6.8) |
where, due to the reciprocity, the point ξξ (corresponding in Eq (6.5) to a source point) has been transformed into a field point outside the domain. By changing the dummy variable (ξξ→y)∈VB and the index α,β,...→a′,b′,... in order to keep the convention that field points are associated with latin lettering, the Green function can be expressed as
Ga α′(x,ζζ)=Sa α′(x−ζζ)+D aa′Sa′ α′(y−ζζ) | (6.9) |
where now
D aa′=1meas(Ω)∫ΩdΩ(y)∞∑n=0Maa′na′(y,x)n!∇a′n,Maa′na′(y,x)=−∫∂Vfga′ a″ | (6.10) |
Although Eqs (6.7) and (6.10) are equivalent, their physical meaning in slightly different. In Eq (6.7), the Green function is expressed as a combination of singular solutions of the unbounded Stokes equation, with poles in \Omega , weighted by the moments that, in turn, depend on the pole {\pmb \zeta} entering the original problem Eq (6.1). Conversely, in Eq (6.10) the field variable enters in the expression of the operator \mathcal{D}_{a'}^{\ a} through the moments, and the regular part, solution of the Stokes equations as a whole, is a combination of terms each of which individually is not a solution of the Stokes equation.
The operators \mathcal{D}^\alpha_{\ \alpha'} defined by Eq (6.6), depend on the pole {\pmb \zeta} via the moments, and consequently, for each {\pmb \zeta} , a new system of moments is defined, determining a different operator \mathcal{D}^\alpha_{\ \alpha'} . For this reason, it is convenient to introduce a new operator, independent of the position of the pole, and such that, its action on the Stokeslet outside the domain of the fluid furnishes the Green function. To this purpose, let us assume that the geometry of the problem is such that there exists a bijective correspondence between points inside {\pmb \zeta} and outside {\pmb \xi} the domain of the fluid, defined by a smooth and invertible function r ,
\begin{equation} {\pmb \xi} = r^{-1}({\pmb \zeta }), \qquad {\pmb \zeta} = r({\pmb \xi}) \end{equation} | (6.11) |
As addressed in Appendix A, and following the Ruse approach to bitensor calculus [25], Eq (6.11) enables us to view {\pmb \xi} and {\pmb \zeta} as conjugate points in two different metric spaces, such that tensorial quantities defined at a point in one of the two spaces can be transported to the conjugate point of the other space via the parallel propagator
\begin{equation} g_{\alpha \beta}({\pmb \xi}) = g^{\ \alpha'}_{\alpha}({\pmb \xi},{\pmb \zeta}) g^{\ \beta'}_{\beta}({\pmb \xi},{\pmb \zeta}) g_{\alpha' \beta'}({\pmb \zeta}) \end{equation} | (6.12) |
where the parallel propagator is given by
\begin{equation} g^{\alpha}_{\ \alpha'}({\pmb \xi},{\pmb \zeta}) = \dfrac{\partial \xi^\alpha}{\partial \zeta^{\alpha'}} \end{equation} | (6.13) |
It follows from Eq (6.12) and from the above bitensorial interpretation of the bijective correspondence Eq (6.11) between point in the flow domain and image points outside it, that the stress tensor \mathrm{\Sigma}_{\alpha'}^{ab}({\bf x}, { \pmb \zeta }) can be parallel transported from the point {\pmb \zeta} to the point {\pmb \xi}
\begin{equation} \mathrm{\Sigma}_{\alpha}^{ab}({\bf x},r({\pmb \xi})) = g^{\ \alpha'}_{\alpha}({\pmb \zeta},{\pmb \xi}) \mathrm{\Sigma}_{\alpha'}^{ab}({\bf x},{\pmb \zeta}) \end{equation} | (6.14) |
Substituting Eq (6.14) into Eq (6.6) one obtains
\begin{equation} M^{\alpha}_{\alpha' {\pmb \alpha}_n}({\pmb \xi},{\pmb \zeta}) = g^{\ \beta}_{\alpha'}({\pmb \xi},{\pmb \zeta}) M^{\alpha}_{\beta {\pmb \alpha}_n}({\pmb \xi},r({\pmb \xi})) \end{equation} | (6.15) |
where
\begin{equation} M^{\alpha}_{\beta {\pmb \alpha}_n}({\pmb \xi},r({\pmb \xi})) = - \int_{\partial V_f} ({\bf x} - {\pmb \xi})_{{\pmb \alpha}_n} g^\alpha_{\ a}({\pmb \xi},{\bf x}) \frac{\mathrm{\Sigma}_{\beta}^{ab}({\bf x},r({\pmb \xi}))}{8\pi} n_b({\bf x}) dS({\bf x}) \end{equation} | (6.16) |
Enforcing Eq (6.15), it is possible to express the operator \mathcal{D}^\alpha_{\ \alpha'} in terms of a reflection operator independent of the pole {\pmb \zeta} , and such that the functional dependence on {\pmb \zeta} is encompassed in the parallel propagator. For highlighting this delicate issue, let us consider the simplest case where \Omega reduces to a point r({\pmb \zeta}) . In this case, it follows from Eq (6.15) that the operator \mathcal{D}^\alpha_{\ \alpha'} attains the form
\begin{equation} \mathcal{D}^\alpha_{\ \alpha'} = g^{\ \beta}_{\alpha'}({\pmb \zeta},{\pmb \xi})\mathcal{R}^\alpha_{\beta}, \qquad \mathcal{R}^\alpha_{\beta} = \sum\limits_{n = 0}^{\infty}\dfrac{ M^{\alpha}_{\beta {\pmb \alpha}_n}({\pmb \xi},r({\pmb \xi}))}{n!}\nabla^{{\pmb \alpha} _n} \end{equation} | (6.17) |
The operator \mathcal{R}^\alpha_{\beta} furnishes the regular part of the Green function starting from the Stokeslet, independently on the source point {\pmb \zeta} and, for the reasons discussed below, it can be referred to as the reflection operator of the bounded flow problem.
By Eqs (6.7) and (6.9), the operator \mathcal{R}^\alpha_{\beta} can be applied on equal footing either at the source or at the field point. In the first case, the Green function reads
\begin{equation} \mathrm{G}^a_{\ \alpha'}({\bf x},{\pmb \zeta}) = \mathsf{S}^a_{\ \alpha'}({\bf x}-{\pmb \zeta}) + g^{\ \beta}_{\alpha'}({\pmb \zeta},{\pmb \xi})\mathcal{R}^\alpha_{\beta}\mathsf{S}^a_{\ \alpha}({\bf x}-{\pmb \xi}) \end{equation} | (6.18) |
In the second case, i.e., by applying the operator at the field point, an alternative representation of the Green function follows
\begin{equation} \mathrm{G}^a_{\ \alpha'}({\bf x},{\pmb \zeta}) = \mathsf{S}^a_{\ \alpha'}({\bf x}-{\pmb \zeta}) + g^{a}_{\ b'}({\bf x},{\bf y})\mathcal{R}^{b'}_{a'}\mathsf{S}^{a'}_{\ \alpha'}({\bf y}-{\pmb \zeta}) \end{equation} | (6.19) |
where {\bf y} = r^{-1}({\bf x}) and {\bf x} = r({\bf y}) . The latter expression permits to interpret the regular part of the Green function as a "reflected field" of the ambient flow, that in the present case is given by a Stokeslet centered at the point {\pmb \zeta} . In fact, the operator \mathcal{R}^{\ b}_{a'} furnishes a continuation of the Stokes solution with homogeneous Dirichlet boundary conditions in the external domain, usually referred as a reflection principle [32,33]. To show this, consider the integral form of a generic solution vanishing at the boundary \partial V_f [40]
\begin{equation} v^a({\bf x}) = -\int_{\partial V_f}\dfrac{\sigma^{\alpha' \beta'}({\pmb \zeta})n_{\beta'}({\pmb \zeta})}{8\pi\mu}\mathsf{S}^a_{\ \alpha'}({\bf x}-{\pmb \zeta})dS({\pmb \zeta})\qquad {\bf x} \in V_f \end{equation} | (6.20) |
and its continuation in \mathbb{R}^3/V_f ,
\begin{equation} v^{a'}({\bf y}) = \int_{\partial V_f}\dfrac{\sigma^{\alpha \beta}({\pmb \xi})n_\beta({\pmb \xi})}{8\pi\mu}\mathsf{S}^{a'}_{\ \alpha}({\bf y}-{\pmb \xi})dS({\pmb \xi})\qquad {\bf y} \in \mathbb{R}^3/V_f \end{equation} | (6.21) |
Since, by the definition of the disturbed field
\begin{equation} g^{\ \beta}_{\alpha'}({\pmb \zeta},{\pmb \xi})\mathcal{R}^\alpha_{\beta}\mathsf{S}^a_{\ \alpha}({\bf x}-{\pmb \xi}) = -\mathsf{S}^a_{\ \alpha'}({\bf x}-{\pmb \zeta}), \qquad {\bf x}\in \partial V_f \end{equation} | (6.22) |
by using the reciprocal identity \mathsf{S}^a_{\ \alpha'}({\bf x}-{\pmb \zeta}) = \mathsf{S}^{\ a}_{\alpha'}({\pmb \zeta}-{\bf x}) and exchanging latin and greek letters, it easy to verify that
\begin{equation} g^{a}_{\ b'} ({\bf x},{\bf y}) \mathcal{R}^{b'}_{a'}\mathsf{S}^{a'}_{\ \alpha'}({\bf y}-{\pmb \zeta}) = - \mathsf{S}^{a}_{\ \alpha'}({\bf x}-{\pmb \zeta}), \qquad {\pmb \zeta} \in \partial V_f \end{equation} | (6.23) |
Therefore, by applying the operator \mathcal{R}^{b'}_{a'} at the field in (6.21) we obtain at the r.h.s of Eq (6.21) the field defined by Eq (6.20) and the reflection formula can be derived
\begin{equation} v^{a}({\bf x}) = g^{a}_{\ b'} ({\bf x},{\bf y})\mathcal{R}^{b'}_{a'} v^{a'}(r({\bf x})) \end{equation} | (6.24) |
The reflection formula in Eq (6.24) requires in principle the estimate of infinite terms as the operator \mathcal{R}^{b'}_{a'} admits in general a series expansion in terms of the countable system of moments. It is known from harmonic function theory, that if the reflection operator (e.g., associated with an electrostatic problem) possesses a finite number of non-vanishing terms, then the boundary is either a plane or a sphere [47] and the relation equivalent to Eq (6.24) is referred as a point-to-point reflection principle. In the case of the solutions of the Stokes problem, that involves biharmonic functions, it is known that a point-to-point reflection principle does not hold even for spherical boundaries, and a weaker point-to-set principle [34,47,48] should be considered, where a bijective relation occurs between a point {\bf x} in the fluid domain and a set parameterized by its conjugate point {\bf y} = r(\bf{x}) in the complementary domain.
Equation (6.17) and the analysis developed in the previous Section indicate the close relation (duality) between the image system of singularities of a bounded flow problem and the formulation of a reflection principle, as the two problems are governed by essentially the same operators \mathcal{D}^\alpha_{\ \alpha'} and {\mathcal R}_\beta^\alpha , parallel transported between a source point and its conjugate image. The duality between image system and reflection principle has been practically neglected in Stokesian hydrodynamics. Several works have investigated the image system of singularities near a plane [36,37] or near spherical boundaries [46], and, almost independently, parallel works on reflected fields near a planar [35] and spherical boundaries [34] has been published. In our opinion, the main difficulty in recognizing a common formal structure underlying image systems and reflection principle in Stokesian hydrodynamics stems from the tensorial nature of the operators involved, and by the need of a parallel transport between conjugate points. The introduction of the bitensorial formalism for hydrodynamic Green functions has made possible to highlight this issue.
The duality between the image system of singularities and the existence of a reflection principle make it possible to transfer and apply methods and techniques developed for solving one of these two problems to the other one. The next Section provides an application of this principle in connection with the problem of singularities bounded by planar boundaries.
Below, the results found in Section 6 are applied to the problem of the singularities of a flow bounded by a rigid plane. In this case, the function r transforming points {\bf x} \in V_f into conjugate points {\bf y} \in \mathbb{R}^3 /V_f is given by the mirror operator {\bf J} = {\bf I} - 2 \, {\bf n} \otimes {\bf n} , {\bf I} being the identity matrix, and {\bf n} the unit normal to the plane, so that {\bf y} = {\bf J}\cdot {\bf x} and {\bf x} = {\bf J}\cdot {\bf y} , since {\bf J}^2 = {\bf I} .
Consider a Cartesian coordinate system ( X_1 , X_2 , X_3 ) with the origin on the plane and such that the flow domain corresponds to X_3 > 0 . Let {\bf x} \in V_f with coordinates ( x_1 , x_2 , x_3 ), and its mirror point {\bf y} \in \mathbb{R}^3/V_f with coordinates y_{a'} = J_{a' a} x_a .
The parallel propagator (see Appendix A) between these conjugate points is given by
\begin{equation} g_{a a'}({\bf x}, {\bf y}) = \dfrac{\partial y^{a'}}{\partial x^a} = J_{a a'} \end{equation} | (7.1) |
The reflection operator acting at the point {\bf y} , corresponding to Eq (6.24) is the so called Lorentz mirror operator [1,34]
\begin{equation} \mathcal{R}_{a' b'} = -J_{a' b'}-2({\bf y} \cdot {\bf n}) \nabla_{a'}\delta_{3 b'}+({\bf y}\cdot {\bf n})^2 \Delta_{x} \delta_{a' b'} \end{equation} | (7.2) |
The Green function of the Stokes flow centered at the source point {\pmb \zeta} \in V_f can be obtained either by applying the reflection operator at the field point, according to Eq (6.19), or at the source point, according to Eq (6.18). In the first case we have
\begin{eqnarray} \mathrm{G}_{a \alpha'} ({\bf x}, {\pmb \zeta}) = \mathsf{S}_{a \alpha'} ({\bf x}-{\pmb \zeta})+J_{a b'} \\ [-J_{b' a'}-2y_3 \nabla_{b'}\delta_{3 a'} +y_3^2\Delta_{x}\delta_{b' a'}] \mathsf{S}_{a' \alpha'} ({\bf y}-{\pmb \zeta}) = \\ \mathsf{S}_{a \alpha'} ({\bf x}- {\pmb \zeta})-\mathsf{S}_{a \alpha'} ({\bf J} \cdot {\bf x}- {\pmb \zeta})+2x_3 J_{a b'}[\nabla_{b'}\mathsf{S}_{3 \alpha'} ({\bf J} \cdot {\bf x}-{\pmb \zeta}) \\ +\dfrac{x_3}{2}\Delta_{x} \mathsf{S}_{b' \alpha'} ({\bf J} \cdot {\bf x}- {\pmb \zeta}) ] \end{eqnarray} | (7.3) |
while the application at the source point provides
\begin{eqnarray} \mathrm{G}_{a\alpha'} ({\bf x}, {\pmb \zeta}) = \mathsf{S}_{a \alpha'} ({\bf x}- {\pmb \zeta})+J_{\alpha' \beta} \\ [-J_{\beta \alpha}-2({\pmb \xi} \cdot {\bf n}) \nabla_{\beta}\delta_{3 \alpha} +({\pmb \xi} \cdot {\bf n})^2\Delta_{\xi} \delta_{\beta \alpha}] \mathsf{S}_{a \alpha} ({\bf x}-{\pmb \xi}) = \\ \mathsf{S}_{a \alpha'} ({\bf x}- {\pmb \zeta})- \mathsf{S}_{a \alpha'} ({\bf x}- {\pmb \xi}) -2({\pmb \xi} \cdot {\bf n}) \\J_{\alpha' \beta}[\nabla_{\beta} \mathsf{S}_{a 3}({\bf x}-{\pmb \xi}) -\dfrac{({\pmb \xi} \cdot {\bf n})}{2} \Delta_{\xi} \mathsf{S}_{a \beta}({\bf x}-{\pmb \xi})] \end{eqnarray} | (7.4) |
with the expression for the pressure
\begin{equation} \mathrm{P}_{\alpha'} ({\bf x}, {\pmb \zeta}) = \mathsf{P}_{\alpha'} ({\bf x}- {\pmb \zeta}) - \mathsf{P}_{\alpha'} ({\bf x}- {\pmb \xi}) -2({\pmb \xi} \cdot {\bf n}) J_{\alpha' \beta}\nabla_{\beta} \mathsf{P}_{3}({\bf x}-{\pmb \xi}) \end{equation} | (7.5) |
and for the stress tensor
\begin{eqnarray} \mathrm{\Sigma}_{a b \alpha'} ({\bf x}, {\pmb \zeta}) = \mathsf{\Sigma}_{a b \alpha'} ({\bf x}- {\pmb \zeta})- \mathsf{\Sigma}_{a b \alpha' } ({\bf x}- {\pmb \xi}) \\ -2({\pmb \xi} \cdot {\bf n}) J_{\alpha' \alpha} [\nabla_{\alpha} \mathsf{\Sigma}_{a b 3}({\bf x}-{\pmb \xi}) -\dfrac{({\pmb \xi} \cdot {\bf n})}{2} \Delta_{\xi} \mathsf{\Sigma}_{a b \alpha}({\bf x}-{\pmb \xi})] \end{eqnarray} | (7.6) |
Since the pole is fixed at {\pmb \xi} = {\bf J} \cdot {\pmb \zeta} = (0, 0, -h) , we obtain the singular form
\begin{equation} \mathrm{G}_{a\alpha'} ({\bf x}, {\pmb \zeta}) = \mathsf{S}_{a \alpha'} ({\bf x}-{\pmb \zeta})- \mathsf{S}_{a \alpha'} ({\bf x}- {\pmb \xi}) +2 h J_{\alpha' \beta}[ \mathsf{S}_{a 3;\beta}({\bf x}-{\pmb \xi}) -h \mathsf{D}_{a \beta}({\bf x}-{\pmb \xi})] \end{equation} | (7.7) |
and
\begin{equation} \mathrm{P}_{\alpha'} ({\bf x}, {\pmb \zeta}) = \mathsf{P}_{\alpha'} ({\bf x}- {\pmb \zeta}) - \mathsf{P}_{\alpha'} ({\bf x}- {\pmb \xi}) +2h J_{\alpha' \beta}\nabla_{\beta} \mathsf{P}_{3}({\bf x}-{\pmb \xi}) \end{equation} | (7.8) |
\begin{equation} \mathrm{\Sigma}_{a b\alpha'} ({\bf x}, {\pmb \zeta}) = \mathsf{\Sigma}_{a b \alpha'} ({\bf x}- {\pmb \zeta})- \mathsf{\Sigma}_{a b \alpha' } ({\bf x}- {\pmb \xi}) \\ +2h J_{\alpha' \alpha} \left[\nabla_{\alpha} \mathsf{\Sigma}_{a b 3 }({\bf x}-{\pmb \xi}) +\dfrac{h}{2} \Delta_{\xi} \mathsf{\Sigma}_{a b \alpha }({\bf x}-{\pmb \xi})\right] \end{equation} | (7.9) |
The singular representation of the Green function Eq (7.7), here obtained simply by applying the Lorentz reflection operator at the source point, coincides with the result obtained by Blake using a much more elaborate approach involving the Fourier-Hankel transforms [36,37].
Blake and Chwang in [36,37] have obtained the singular reflection systems related to bounded Stokeslet, Sourcelet and Rotlet by applying the Fourier-Hankel transforms to separate and distict problems specified by the boundary conditions adopted. In point of fact, the operator formalism developed in Section 6 permits to obtain any higher-order singularity in a unitary way, by simply differentiating the Green's function at the pole, Eqs (7.4)–(7.6).
To begin with, consider the bounded Source Dipole \mathrm{D}_{a\alpha'}({\bf x}, {\pmb \zeta}) defined by (3.24), applying the Laplacian operator -\Delta_{\zeta}/2 to the expression (7.4). Being the Laplacian operator invariant with respect to any Euclidean transformation, and thus under the reflection transformation {\pmb \zeta} = {\bf J}\cdot {\pmb \xi} , we have \Delta_{\xi} = \Delta_{\zeta} , and therefore
\begin{equation} \mathrm{D}_{a\alpha'} ({\bf x},{\pmb \zeta}) = \mathsf{D}_{a \alpha'} ({\bf x}- {\pmb \zeta})- \mathsf{D}_{a \alpha'} ({\bf x}-{\pmb \xi}) +J_{\alpha' \beta}\Delta_{\xi}[({\pmb \xi} \cdot {\bf n}) \nabla_{\beta} \mathsf{S}_{a 3}({\bf x}-{\pmb \xi})] \\ +J_{\alpha' \beta}\Delta_{\xi}[({\pmb \xi} \cdot {\bf n})^2 \mathsf{D}_{a \beta}({\bf x}-{\pmb \xi})] \end{equation} | (7.10) |
where, enforcing the identity,
\begin{equation} \Delta_{\xi}\mathsf{S}_{a3;\beta }({\bf x}-{\pmb \xi}) = \nabla_\beta\Delta_{\xi}\mathsf{S}_{a3}({\bf x}-{\pmb \xi}) = -2\mathsf{D}_{a3;\beta}({\bf x}-{\pmb \xi}) \end{equation} | (7.11) |
the third term at the r.h.s of Eq (7.10) reads
\begin{equation} J_{\alpha' \beta}\Delta_{\xi}[({\pmb \xi} \cdot {\bf n}) \nabla_{\beta} \mathsf{S}_{a 3}({\bf x}-{\pmb \xi})] = J_{\alpha' \beta}\nabla_\gamma(\delta_{\gamma 3}\mathsf{S}_{a3;\beta }({\bf x}-{\pmb \xi})\\+\xi_3\mathsf{S}_{a3;\beta \gamma}({\bf x}-{\pmb \xi})) = J_{\alpha' \beta}(2\mathsf{S}_{a3;\beta3 }({\bf x}-{\pmb \xi})-2\xi_3\mathsf{D}_{a3;\beta }({\bf x}-{\pmb \xi})) \end{equation} | (7.12) |
The fourth term in Eq (7.10) can be simplified as
\begin{equation} J_{\alpha' \beta}\Delta_{\xi}[({\pmb \xi} \cdot {\bf n})^2 \mathsf{D}_{a \beta}({\bf x}-{\pmb \xi})] = J_{\alpha' \beta}\nabla_\gamma( 2\delta_{3\gamma}\xi_3 \mathsf{D}_{a \beta}({\bf x}-{\pmb \xi})+ \xi_3 ^2 \mathsf{D}_{a \beta;\gamma} ({\bf x}-{\pmb \xi})) = \\J_{\alpha' \beta}( 2\mathsf{D}_{a \beta}({\bf x}-{\pmb \xi})+4\xi_3 \mathsf{D}_{a \beta; 3}({\bf x}-{\pmb \xi})) \end{equation} | (7.13) |
so that the singular representation of the Source Dipole reads
\begin{equation} \mathrm{D}_{a \alpha'}({\bf x}, {\pmb \zeta}) = \mathsf{D}_{a \alpha'}({\bf x}-{\pmb \zeta}) -\mathsf{D}_{a \alpha'}({\bf x}-{\pmb \xi}) +2J_{\alpha' \beta} \\ \bigg( \mathsf{D}_{a \beta}({\bf x}-{\pmb \xi}) +\mathsf{S}_{a3;\beta 3 }({\bf x}-{\pmb \xi}) +\xi_3 \mathsf{D}_{a3;\beta}({\bf x}-{\pmb \xi}) \bigg) \end{equation} | (7.14) |
and since \xi_3 = -h , Eq (7.14) becomes
\begin{equation} \mathrm{D}_{a \alpha'}({\bf x}, {\pmb \zeta}) = \mathsf{D}_{a \alpha'}({\bf x}- {\pmb \zeta}) -\mathsf{D}_{a \alpha'}({\bf x}-{\pmb \xi}) +2J_{\alpha' \beta} \\\bigg( \mathsf{D}_{a \beta}({\bf x}-{\pmb \xi}) +\mathsf{S}_{a3;\beta 3 }({\bf x}-{\pmb \xi}) -h \mathsf{D}_{a3;\beta}({\bf x}-{\pmb \xi}) \bigg) \end{equation} | (7.15) |
The associated pressure field can obtained by applying the same operator -\Delta_{\zeta}/2 = -\Delta_{\xi}/2 to the pressure Green function Eq (7.5). Since the unbounded pressure field is a potential vector field with respect to the source point coordinates, the only non vanishing contribution is given by the third term at the r.h.s of Eq (7.5), and therefore
\begin{eqnarray} -\dfrac{\Delta_{\zeta}\mathrm{P}_{\alpha'} ({\bf x}, {\pmb \zeta})}{2} = 4\pi(\nabla_{\alpha'}\delta({\bf x}-{\pmb \zeta})- \delta_{\alpha'\alpha}\nabla_{\alpha}\delta({\bf x}-{\pmb \xi})) -J_{\alpha' \beta}\Delta_{\xi}\left(\xi_3 \nabla_{\beta}\mathsf{P}_{3}({\bf x}-{\pmb \xi})\right) = \\ 4\pi(\nabla_{\alpha'}\delta({\bf x}-{\pmb \zeta})- \delta_{\alpha'\alpha}\nabla_{\alpha}\delta({\bf x}-{\pmb \xi})) -2J_{\alpha' \beta}\delta_{3\gamma}\nabla_{\beta}\nabla_{\gamma}\mathsf{P}_{3}({\bf x}-{\pmb \xi}) \end{eqnarray} | (7.16) |
Figure 2 provides the schematic representation of the unbounded singularities at the image pole necessary to cancel the velocity field at the plane due to the unbounded Source Doubled at the pole in the fluid domain. Panel (a) refers to \mathrm{D}_{a1} = D_{a2} , panel (b) to D_{a3} . The vector plot of the bounded Source Dipole defined by Eq (7.15) is depicted in Figure 3.
In the far field, |{\bf x}| > > h , we have
\begin{equation} \mathrm{D}_{a \alpha'}({\bf x},{\pmb \zeta}) = 2J_{\alpha' \alpha} \bigg( \mathsf{D}_{a \alpha}({\bf x})+ \mathsf{S}_{3 \alpha;3 a}({\bf x}) \bigg)+o(1/|{\bf x}|^3), \qquad |{\bf x}|\gg |{\pmb \zeta}| \end{equation} | (7.17) |
To obtain the Stokes Doublet Eq (3.16), we can apply the covariant derivative at the pole of the Green function. The bounded solution of the Rotlet (3.22) giving the antisymmetric part of the Stokes doublet can be found, according to Eq (3.19), by applying the curl at the pole of the Green function to obtain
\begin{equation} \mathrm{\Omega}_{a \gamma'}({\bf x}, {\pmb \zeta}) = \mathsf{\Omega}_{a \gamma'}({\bf x}-{\pmb \zeta}) -\mathsf{\Omega}_{a \gamma'}({\bf x}-{\pmb \xi}) +2\epsilon_{\beta' \gamma' 3}\bigg( \mathsf{E}_{ a 3 \beta'}({\bf x}-{\pmb \xi} ) +\xi_3 \mathsf{D}_{a \beta'}({\bf x}-{\pmb \xi} ) \bigg) \end{equation} | (7.18) |
and the associated pressure reads
\begin{equation} \dfrac{\varepsilon_{\gamma'\epsilon'\eta'}\nabla_{\eta'}\mathrm{P}_{\epsilon'}({\bf x},{\pmb \zeta})}{2} = \varepsilon_{\gamma'\epsilon'\eta'}\nabla_{\eta'}\xi_3 J_{\epsilon' \beta}\nabla_{\beta}\mathsf{P}_{3}({\bf x}-{\pmb \xi}) = \varepsilon_{\gamma'\epsilon'3}\delta_{\epsilon'\beta}\nabla_{3}\nabla_{\beta}\mathsf{P}_{3}({\bf x}-{\pmb \xi}) \end{equation} | (7.19) |
In the far field we have the asymptotic scaling
\begin{equation} \mathrm{\Omega}_{a \alpha'}({\bf x},{\pmb \zeta} ) = 2 \epsilon_{\beta' \alpha' 3} \mathsf{T}_{ a 3 \beta'}({\bf x})(1-\delta_{\alpha 3}) +o(1/|{\bf x}|^2), \qquad |{\bf x}|\gg |{\pmb \zeta}| \end{equation} | (7.20) |
The vector plot of the Rotlet is depicted in Figure 4.
To obtain the bounded Strainlet Eq (3.20), i.e., the symmetric part of the Stokes Doublet, we could evaluate (\nabla_{\beta'} \mathrm{G}_{a\alpha'} +\nabla_{\alpha'}\mathrm{G}_{a\beta'})/2 . Alternatively, it is more convenient to use Eq (3.21), substituting in it Eq (7.5) for the pressure, and Eq (7.6) for the stress tensor of the bounded Green function
\begin{equation} \mathrm{E}_{a \alpha' \beta'}({\bf x},{\pmb \zeta}) = \dfrac{\delta_{\alpha' \beta'}}{2} \mathrm{P}_{a}({\pmb \zeta},{\bf x}) -\dfrac{1}{2}\mathrm{\Sigma}_{\alpha' \beta' a }({\pmb \zeta},{\bf x}) \end{equation} | (7.21) |
where
\begin{equation} \mathrm{P}_{a} ( {\pmb \zeta},{\bf x}) = \mathsf{P}_{a} ({\pmb \zeta}-{\bf x}) - \mathsf{P}_{a} ({\pmb \zeta}-{\bf y}) -2({\bf y} \cdot {\bf n}) J_{a a'}\nabla_{a'} \mathsf{P}_{3}({\pmb \zeta}-{\bf y}) \end{equation} | (7.22) |
and
\begin{eqnarray} \mathrm{\Sigma}_{\alpha'\beta' a } ({\pmb \zeta},{\bf x}) = \mathsf{\Sigma}_{\alpha'\beta' a} ({\pmb \zeta}-{\bf x})- \mathsf{\Sigma}_{\alpha' \beta' a} ({\pmb \zeta} -{\bf y}) -2({\bf y} \cdot {\bf n}) J_{a a'} \\ [\nabla_{a'} \mathsf{\Sigma}_{\alpha' \beta' 3 }({\pmb \zeta}-{\bf y}) -\dfrac{({\bf y} \cdot {\bf n})}{2} \Delta_{y} \mathsf{\Sigma}_{\alpha' \beta' a'}({\pmb \zeta}-{\bf y})] \end{eqnarray} | (7.23) |
Since \Delta_{\xi} = \Delta_{\zeta} = \Delta_{x} = \Delta_{y} , in this particular case, where the boundary of the fluid is a plane, it is possible to define the bounded Source \mathrm{M}_{a}({\bf x}, {\pmb \zeta}) = -\mathrm{P}_{a}({\pmb \zeta}, {\bf x}) /2 and the bounded Stresslet \mathrm{T}_{a \alpha' \beta'}({\bf x}, {\pmb \zeta}) = -\mathrm{\Sigma}_{\alpha' \beta' a }({\pmb \zeta}, {\bf x})/2 , that are the bounded counterparts of the Sourcelet defined in Eq (3.32) and the Stresslet in Eq (3.33). Therefore, the Strainlet can be expressed as
\begin{equation} \mathrm{E}_{a \alpha' \beta'}({\bf x},{\pmb \zeta}) = -\delta_{\alpha' \beta'} \mathrm{M}_{a}({\bf x},{\pmb \zeta}) +\mathrm{T}_{a \alpha' \beta'}({\bf x},{\pmb \zeta}) \end{equation} | (7.24) |
By making the following transformation
\begin{eqnarray} {\bf y}\cdot {\bf n} = y_3 = ({\bf y} - {\pmb \xi})_3+\xi_3 \end{eqnarray} | (7.25) |
we obtain for \mathrm{M}_{a}({\bf x}, {\pmb \zeta}) and \mathrm{T}_{a \alpha' \beta'}({\bf x}, {\pmb \zeta}) the following expressions
\begin{equation} \mathrm{M}_{a }({\bf x}, {\pmb \zeta}) = \mathsf{M}_{a}({\bf x}-{\pmb \zeta}) -\mathsf{M}_{a}({\bf x}-{\pmb \xi}) +2 \bigg( \mathsf{T}_{a33}({\bf x}-{\pmb \xi}) +\xi_3 \mathsf{D}_{a3}({\bf x}-{\pmb \xi}) \bigg) \end{equation} | (7.26) |
\begin{eqnarray} \mathrm{T}_{a \alpha'\beta'}({\bf x}, {\pmb \zeta}) = \mathsf{T}_{a \alpha' \beta'}({\bf x}-{\pmb \zeta}) \\+2 \delta_{\alpha'\beta'} \mathsf{T}_{a33}({\bf x}- {\pmb \xi}) -J_{\alpha' \alpha}J_{\beta' \beta} \mathsf{T}_{a \alpha \beta}({\bf x}- {\pmb \xi})+ \\ +2J_{\alpha' \alpha}J_{\beta' \beta}\xi_3 \bigg( -\xi_3 \mathsf{D}_{a \alpha ;\beta}({\bf x}- {\pmb \xi}) -\mathsf{S}_{ a 3;\alpha \beta}({\bf x}- {\pmb \xi})\\ -\delta_{3 \beta} \mathsf{D}_{a \alpha}({\bf x}- {\pmb \xi}) -\delta_{3 \alpha} \mathsf{D}_{a \beta}({\bf x}-{\pmb \xi}) +\delta_{\alpha \beta} \mathsf{D}_{3 a}({\bf x}- {\pmb \xi}) \bigg) \end{eqnarray} | (7.27) |
which possess the following far-field asymptotics
\begin{equation} \mathrm{M}_{a}({\bf x},{\pmb \zeta}) = 2 \mathsf{T}_{ a 3 3}({\bf x}) +o(1/|{\bf x}|^2), \qquad |{\bf x}| \gg |{\pmb \zeta}| \end{equation} | (7.28) |
\begin{equation} \mathrm{T}_{a \alpha'\beta'}({\bf x},{\pmb \zeta}) = 2\mathsf{T}_{a \alpha \beta}({\bf x})(1-\delta_{\alpha \beta}-\delta_{\alpha 1}\delta_{\beta 2}-\delta_{\alpha 2}\delta_{\beta 1})\\+2\mathsf{T}_{a33}({\bf x})\delta_{\alpha \beta}+o(1/|{\bf x}|^2) , \qquad |{\bf x}| \gg |{\pmb \zeta}| \end{equation} | (7.29) |
Gathering Eqs (7.26) and (7.27) and substituting them into Eq (7.24), the analytic expression for the bounded Strainlet follows
\begin{eqnarray} \mathrm{E}_{a \alpha'\beta'}({\bf x}, {\pmb \zeta}) = \mathsf{E}_{a \alpha' \beta'}({\bf x}-{\pmb \zeta})\\ -J_{\alpha' \alpha}J_{\beta' \beta} \mathsf{E}_{a \alpha \beta}({\bf x}- {\pmb \xi})+ \\ +2J_{\alpha' \alpha}J_{\beta' \beta}\xi_3 \bigg( -\xi_3 \mathsf{D}_{a \alpha ;\beta}({\bf x}- {\pmb \xi}) -\mathsf{S}_{ a 3;\alpha \beta}({\bf x}- {\pmb \xi}) \\-\delta_{3 \beta} \mathsf{D}_{a \alpha}({\bf x}- {\pmb \xi}) -\delta_{3 \alpha} \mathsf{D}_{a \beta}({\bf x}-{\pmb \xi}) \bigg) \end{eqnarray} | (7.30) |
and putting \xi_3 = -h , one obtains
\begin{eqnarray} \mathrm{E}_{a \alpha'\beta'}({\bf x}, {\pmb \zeta}) = \mathsf{E}_{a \alpha' \beta'}({\bf x}-{\pmb \zeta}) -J_{\alpha' \alpha}J_{\beta' \beta} \mathsf{E}_{a \alpha \beta}({\bf x}- {\pmb \xi})+ \\ -2hJ_{\alpha' \alpha}J_{\beta' \beta} \bigg( h \mathsf{D}_{a \alpha ;\beta}({\bf x}- {\pmb \xi}) -\mathsf{S}_{ a 3;\alpha \beta}({\bf x}- {\pmb \xi}) -\delta_{3 \beta}\\ \mathsf{D}_{a \alpha}({\bf x}- {\pmb \xi}) -\delta_{3 \alpha} \mathsf{D}_{a \beta}({\bf x}-{\pmb \xi}) \bigg) \end{eqnarray} | (7.31) |
The vector plot of the bounded Strainlet is depicted in Figure 5.
In this article we have applied the bitensor calculus to the singularity method in Stokes flow. The nature of the bitensorial formalism, that distinguishes between source and field points of singular fields, allows us to manipulate mathematically singularities without ambiguity, resulting useful even if the fluid domain is regarded as a flat space.
We have provided a clear definition of the singularities in Stokes flow, specifying the associated non-homogeneous equations and boundary conditions, obtaining the most common unbounded singularities as a particular case of the more general bounded counterparts. Although this topic can be found in any monograph on Stokesian hydrodynamics [3,4], the detailed description of some of the most common Stokesian singularities has never been, to the best of our knowledge, addressed in the hydrodynamic literature.
Moreover, an explicit formulation of the singular method, providing a way for expressing bounded flows in terms of unbounded singularities, has been derived. To this aim, we have defined the moments as volume integrals in the domain of the obstacle, and we have used a tensorial moment theory to obtain the integro-differential operator yielding the disturbance flow once applied at the pole of the Stokeslet in term of a countable set of moments. As shown in Appendix B, this definition is coincident whit the surface moments defined by Ichiki [31] for no-slip spheres immersed in a Stokes flow. To evaluate the moments of a specific flow problem, we developed a method based on geometrical moments related to the obstacle immersed in purely n -th order ambient flow. This method is useful either in numerical applications or in theoretical analysis. We have shown that starting from the geometrical moments it is possible to define a n -th order Faxén operator, i.e., an operator satisfying a generalized Faxén theorem. In addition, we found that a generic disturbance flow can be developed in a series of Faxén operators applied at the Stokeslet's pole, thus expressing a generic field as a series of simpler disturbance flows associated with purely n -th order ambient flows.
Enforcing the reciprocal symmetry of the Green function, we have shown that it is possible to apply the operator both to the source and and to the field point of the Stokeslet, in order to obtain the disturbance contribution to the flow field. The main consequence of the latter result is that, whenever it is possible to define a reflection operator, this operator coincides with the operator derived from moment theory furnishing the image system of singularity. This result, applied to the Green function bounded by a plane provides an alternative way for expressing the hydrodynamic singularities which is simpler than the method used by Blake [36] involving Fourier-Hankel transforms, and it has been used to derive other singularities, such as the Source Dipole and the Strainlet.
The practical application of the theory to specific bounded Stokes problem will be addressed in forthcoming works.
The authors declare no conflict of interest.
The development of bitensor calculus has followed two parallel pathways: an algebraic [25] and purely geometric approach [26,27]. As the algebraic approach is particularly relevant in the present hydrodynamic theory of bounded Green functions, and moreover it is scarsely mentioned in the literature, this brief review on bitensor calculus is mainly focused on this formulation, addressing its connection with the geometric theory at the end of this Appendix.
In [25], Ruse defines bitensors as follows: Let {\bf x} = (x^1, ..., x^n) and {\pmb \xi} = (\xi^1, ..., \xi^m) be two set of independent variables and let f^{a'}({\bf x}) , a' = 1, \dots, n be n functions dependent on {\bf x} and \phi^{\alpha'}({\pmb \xi}) , \alpha' = 1, \dots, m , m functions dependent on {\pmb \xi} , such that we can define the new variables
\begin{equation} \nonumber x^{b'} = f^{b'}({\bf x}) \, , \;\; b' = 1,\dots,n\,, \qquad \xi^{\beta'} = \phi^{\beta'}({\pmb \xi}) \, , \;\; \beta' = 1,\dots,m \label{A1} \end{equation} |
Let T^{a \alpha}({\bf x}, {\pmb \xi}) denote the array of n \times m functions depending on both {x}^a and {\xi}^\alpha
\begin{equation} \left(\begin{array}{ccc} T^{11}({\bf x}, {\pmb \xi}) & ... & T^{1 m}({\bf x}, {\pmb \xi})\\... & ... & ...\\ T^{n 1}({\bf x}, {\pmb \xi}) & ... & T^{n m}({\bf x}, {\pmb \xi}) \end{array} \right) \end{equation} | (A.1) |
Moreover, let T^{a' \alpha}(\bf x', {\pmb \xi}) be a set of functions depending on the variables {\bf x}' and {\pmb \xi} , T^{a \alpha'}({\bf x}, {\pmb \xi}') a set of functions depending on the variables {\bf x} and {\pmb \xi}' , T^{a' \alpha'}({\bf x}', {\pmb \xi}') a set of functions depending on the variables {\bf x}' and {\pmb \xi}' . If these functions are related by the equations
\begin{equation} {T}^{a \alpha}({\bf x}', {\pmb \xi}) = T^{b \alpha}({\bf x}, {\pmb \xi}) \, \dfrac{\partial {x}^{a'}}{\partial x^b} \end{equation} | (A.2) |
\begin{equation} T^{a \alpha'}({\bf x}, {\pmb \xi}') = T^{a \beta}({\bf x}, {\pmb \xi}) \, \dfrac{\partial {\xi}^{\alpha'}}{\partial \xi^\beta} \end{equation} | (A.3) |
\begin{equation} T^{a' \alpha'}({\bf x}', {\pmb \xi}') = T^{b \beta}({\bf x}, {\pmb \xi}) \, \dfrac{\partial {x}^{a'}}{\partial x^b}\dfrac{\partial {\xi}^{\alpha'}}{\partial \xi^\beta} \end{equation} | (A.4) |
then they are the components of the bivector {\bf T} expressed in the systems of coordinates ({\bf x}, {\pmb \xi}) , ({\bf x}', {\pmb \xi}) , ({\bf x}, {\pmb \xi}') , ({\bf x}', {\pmb \xi}') , respectively. More generally a set of n^{r+s} \times m^{p+q} functions are the components of a bitensor {\bf T} , if they are related by the equations
\begin{equation} {T}^{a_1'... a_r' \alpha_1... \alpha_q}_{b_1'... b_s' \beta_1... \beta_p}({\bf x}', {\pmb \xi}) = {T}^{c_1... c_r \alpha_1... \alpha_q}_{d_1... d_s \beta_1... \beta_p} ({\bf x}, {\pmb \xi}) \, \dfrac{\partial {x}^{a_1'}}{\partial x^{c_1}}... \dfrac{\partial {x}^{a_r'}}{\partial x^{c_r}} \dfrac{\partial {x}^{d_1}}{\partial {x}^{b_1'}}... \dfrac{\partial {x}^{d_s}}{\partial {x}^{b_s'}} \end{equation} | (A.5) |
\begin{equation} {T}^{a_1... a_r \alpha_1'... \alpha_q'}_{b_1... b_s \beta_1'... \beta_p'}({\bf x}, {\pmb \xi}') = {T}^{a_1... a_r \gamma_1... \gamma_q}_{b_1... b_s \delta_1... \delta_p} ({\bf x}, {\pmb \xi}) \dfrac{\partial {\xi}^{\alpha_1'}}{\partial \xi^{\gamma_1}}... \dfrac{\partial {\xi}^{\alpha_q'}}{\partial {\xi}^{\gamma_q'}} \dfrac{\partial {\xi}^{\delta_1}}{\partial {\xi}^{\beta_1'}}... \dfrac{\partial {\xi}^{\delta_p}}{\partial {\xi}^{\beta_p'}} \end{equation} | (A.6) |
\begin{equation} {T}^{a_1'... a_r' \alpha_1'... \alpha_q'}_{b_1'... b_s' \beta_1'... \beta_p'}({\bf x}', {\pmb \xi}') = {T}^{c_1... c_r \gamma_1... \gamma_q}_{d_1... d_s \delta_1... \delta_p} ({\bf x}, {\pmb \xi}) \dfrac{\partial {x}^{a_1'}}{\partial x^{c_1}}... \\ \dfrac{\partial {x}^{a_r'}}{\partial x^{c_r}} \dfrac{\partial {x}^{d_1}}{\partial {x}^{b_1'}}... \dfrac{\partial {x}^{d_s}}{\partial {x}^{b_s'}} \dfrac{\partial {\xi}^{\alpha_1'}}{\partial \xi^{\gamma_1}}... \dfrac{\partial {\xi}^{\alpha_q'}}{\partial {\xi}^{\gamma_q}} \dfrac{\partial {\xi}^{\delta_1}}{\partial {\xi}^{\beta_1'}}... \dfrac{\partial {\xi}^{\delta_p}}{\partial {\xi}^{\beta_p'}} \end{equation} | (A.7) |
If {\pmb \xi} is kept fixed, then T^{a 1}... T^{a m} are the components ( a = 1, \dots, n ) of m ordinary vectors at {\bf x} , whereas if {\bf x} is kept fixed T^{1 \alpha}... T^{n \alpha} are the components ( \alpha = 1, \dots, m ) of n vectors at {\pmb \xi} . The bitensor T^{a \alpha} is, then, named vector-vector bitensor and, more generally, the bitensor {T}^{a_1... a_r \alpha_1... \alpha_q}_{b_1... b_s \beta_1... \beta_p} is named (r+s) tensor- (p+q) tensor.
Next consider two symmetric scalar- (2) tensor g_{a b}({\bf x}, {\pmb \xi}) and \gamma_{\alpha \beta}({\bf x}, {\pmb \xi}) , and suppose that {\bf x} and {\pmb \xi} are two systems of coordinates of two distinct Riemannian spaces defined respectively by the two metric forms
\begin{eqnarray} ds^2 = g_{ab}({\bf x}, {\pmb \xi}) \, dx^a dx^b \end{eqnarray} | (A.8) |
\begin{eqnarray} d\sigma^2 = \gamma_{\alpha \beta}({\bf x}, {\pmb \xi}) \, d\xi^\alpha d\xi^\beta \end{eqnarray} | (A.9) |
Equations (A.8)–(A.9) define a multiple-infinite set of Riemannian spaces. In fact, fixed the set of variables {\pmb \xi} , Eq (A.8) determines a Riemannian space, while fixing {\bf x} , a Riemannian space is determined by Eq (A.9). In the case that n = m , it is possible to define a vector-vector bitensor k^a_{\ \alpha}({\bf x}, {\pmb \xi}) , belonging to both spaces, so that
\begin{equation} g_{a b}({\bf x}, {\pmb \xi}) = k_a^{\ \alpha}({\bf x}, {\pmb \xi}) \, k_b^{\ \beta}({\bf x}, {\pmb \xi}) \, \gamma_{\alpha \beta}({\bf x}, {\pmb \xi}) \end{equation} | (A.10) |
which represents a system of n(n+1)/2 equations for the n^2 unknown components of k^a_{\ \alpha} (due to the symmetry of g_{ab} and \gamma_{\alpha \beta} ).
Note that, keeping either {\bf x} or {\pmb \xi} fixed, the n ordinary vectors k^1_{\ \alpha}... k^n_{\ \alpha} and k^a_{\ 1}... k^a_{\ n} are orthogonal to each other
\begin{equation} k^c_{\ \beta}k_b^{\ \beta} = \delta^c_b \end{equation} | (A.11) |
and similarly
\begin{equation} k^b_{\ \gamma}k_b^{\ \beta} = \delta^\beta_\gamma \end{equation} | (A.12) |
A particular case occurs when g_{a b}({\bf x}, {\pmb \xi}) = g_{a b}({\bf x}) does not depend on {\pmb \xi} and moreover \gamma_{\alpha, \beta}({\bf x}, {\pmb \xi}) = g_{\alpha \beta}({\pmb \xi}) . In this case, the two metric spaces defined by Eqs (A.8)–(A.9) represent the same metric space at two different points, and Eq (A.10) becomes
\begin{equation} g_{a b}({\bf x}) = k_a^{\ \alpha}({\bf x}, {\pmb \xi}) \, k_b^{\ \beta}({\bf x}, {\pmb \xi}) \, g_{\alpha \beta}({\pmb \xi}) \end{equation} | (A.13) |
In Euclidean spaces, it is always possible to express the component of the metric tensors g_{a b}({\bf x}) , g_{\alpha \beta}({\pmb \xi}) in the same Cartesian coordinate system X_{(i)} = (X_{(1)}, X_{(2)}, X_{(3)}) , (i) = 1, 2, 3 being the indices for the Cartesian components. Thus, from Eq (A.13) one has
\begin{equation} I_{(ij)}\dfrac{\partial X^{(i)}}{\partial x^a}\dfrac{\partial X^{(j)}}{\partial x^b} = k_a^{\ \alpha}({\bf x}, {\pmb \xi})k_b^{\ \beta}({\bf x}, {\pmb \xi})\dfrac{\partial X^{(i)}}{\partial \xi^\alpha}\dfrac{\partial X^{(j)}}{\partial \xi^\beta}I_{(ij)} \end{equation} | (A.14) |
where I_{(ij)} = \rm{diag}(1, 1, 1) from which it follows that
\begin{equation} k_a^{\ \alpha}({\bf x}, {\pmb \xi}) = \dfrac{\partial X^{(i)}}{\partial x^a}\dfrac{\partial \xi^\alpha}{\partial X^{(i)}} \end{equation} | (A.15) |
that reduces to k_a^{\ \alpha}({\bf x}, {\pmb \xi}) = \delta^a_\alpha if we choose the same coordinate system at both points.
If the coordinates of the two points are related by a bijective transformation
\begin{equation} \xi^\alpha = f^\alpha({\bf x}); \qquad x^a = g^\alpha({\pmb \xi}) \end{equation} | (A.16) |
we can consider \xi^\alpha and x^a as two set of coordinates of the same point and, by classical tensor calculus, we have
\begin{equation} g_{a b}({\bf x}) = \dfrac{\partial \xi^\alpha}{\partial x^a}\dfrac{\partial \xi^\beta}{\partial x^b} \, g_{\alpha \beta}({\pmb \xi}) \end{equation} | (A.17) |
Comparing Eq (A.17) with Eq (A.13), we find
\begin{equation} k^a_{\ \alpha}(x^a, \xi^\alpha) = \dfrac{\partial \xi^\alpha}{\partial x^a} \end{equation} | (A.18) |
therefore, an ordinary transformation in the classical tensor calculus, can be viewed as a transformation between two metric spaces
\begin{eqnarray} ds^2 = g_{ab}(g^a({\pmb \xi})) \, dx^a \, dx^b \end{eqnarray} | (A.19) |
\begin{eqnarray} d\sigma^2 = g_{\alpha \beta}(f^\alpha({\bf x})) \, d\xi^\alpha \, d\xi^\beta \end{eqnarray} | (A.20) |
If the two points belong to a generic Riemannian space, it is not always possible to express the components in the same Cartesian coordinate system. However, we can define, at one point, say {\bf x} , a triad of vector e_{(i)}^a({{\bf x}}) , forming locally an orthonormal basis, so that [38,39]
\begin{equation} g_{ab}({{\bf x}})e_{(i)}^a({\bf x}) e_{(j)}^b({\bf x}) = I_{(ij)}; \qquad e_{(i)}^a({\bf x})e^{(i)}_b({\bf x}) = \delta^a_b \end{equation} | (A.21) |
Parallel transporting the vectors e_{(i)}^a({\bf x}) from the point {\bf x} to the point {\pmb \xi} , i.e., integrating the differential equation
\begin{equation} \dfrac{D e_{(i)}^a({\bf z})}{D u} = \dfrac{\partial e_{(i)}^a({\bf z})}{\partial z^k}\dfrac{d z^k}{d u}+\Gamma^{a}_{b k}e_{(i)}^a({\bf z})\dfrac{d z^k}{d u} = 0 \end{equation} | (A.22) |
along the geodetics connecting the point {\bf x} to {\pmb \xi} , {\bf z}(u) being a generic point on the geodetics identified by the parameter u and such that {\bf z}(0) = {\bf x} , we obtain the triad of vectors at the point {\pmb \xi} , that is still orthonormal. Thus,
\begin{equation} g_{\alpha \beta}({\pmb \xi}) \, e_{(i)}^\alpha({\pmb \xi}) \, e_{(j)}^\beta({\pmb \xi}) = I_{(ij)} \, , \qquad e_{(i)}^\alpha({\pmb \xi}) \, e^{(i)}_\beta({\pmb \xi}) = \delta^\alpha_\beta \end{equation} | (A.23) |
Using Eq (A.21) and (A.23), it is possible to express both the metric tensors g_{ab}({\bf x}) and g_{\alpha \beta}({\pmb \xi}) in a common orthonormal basis. Equation (A.14) thus, becomes
\begin{equation} I_{(ij)} \, e^{(i)}_a({\bf x}) \, e^{(j)}_b({\bf x}) = k_a^{\ \alpha}({\bf x}, {\pmb \xi}) \, k_b^{\ \beta}({\bf x}, {\pmb \xi}) e^{(i)}_\alpha({\pmb \xi})\, e^{(j)}_\beta ({\pmb \xi}) \, I_{(ij)} \end{equation} | (A.24) |
obtaining the more general expression for the parallel propagator
\begin{equation} k_a^{\ \alpha}({\bf x}, {\pmb \xi}) = e^{(i)}_a({\bf x})e_{(i)}^\alpha({\pmb \xi}) = k_{\ a}^{\alpha}({\pmb \xi}, {\bf x}) \end{equation} | (A.25) |
In the case the two points {\bf x} and {\pmb \xi} become coincident, we have
\begin{equation} \lim\limits_{{\pmb \xi} \rightarrow {\bf x}} {\bf k}({\bf x}, {\pmb \xi}) = \lim\limits_{{\pmb \xi} \rightarrow {\bf x}} [k_a^{\ \alpha}({\bf x}, {\pmb \xi}) ] = \lim\limits_{{\pmb \xi} \rightarrow {\bf x}} [ e^{(i)}_a({\bf x}) \, e_{(i)}^\alpha({\pmb \xi})] = [\delta^\alpha_a] = {\bf I} \end{equation} | (A.26) |
where [ \cdot ] indicate the whole tensorial entity. Consequently,
\begin{equation} \lim\limits_{{\pmb \xi} \rightarrow {\bf x}} k_{a \alpha}(x^a, \xi^\alpha) = \lim\limits_{{\pmb \xi} \rightarrow {\bf x}} \, g_{ab}({\bf x}) \, k^b_{\ \alpha}({\bf x}, {\pmb \xi}) = g_{ab}({\bf x}) \, , \qquad \lim\limits_{{\bf x} \rightarrow {\pmb \xi}} k_{a \alpha}({\bf x}, {\pmb \xi}) = g_{\alpha\beta}({\pmb \xi}) \end{equation} | (A.27) |
Therefore, it is customary to use the same symbol for indicating either the parallel propagator or the metric tensor
\begin{equation} \nonumber g_a^{\ \alpha}({\bf x},{\pmb \xi}) = k_a^{\ \alpha}({\bf x},{\pmb \xi}) \end{equation} |
To make an example, consider a unit vector p^a({\bf x}) at {\bf x} ,
\begin{equation} g_{ab}({\bf x})p^a({\bf x}) p^b({\bf x}) = 1 \end{equation} | (A.28) |
Using the definition of the parallel propagator Eq (A.13)
\begin{equation} g_{\alpha \beta}({\pmb \xi}) \, k_a^{\ \alpha}({\bf x}, {\pmb \xi}) \, p^a({\bf x}) \, k_b^{\ \beta}({\bf x}, {\pmb \xi}) \, p^b({\bf x}) = g_{\alpha \beta}({\pmb \xi}) \, p^\alpha({\pmb \xi}) \, p^\beta({\pmb \xi}) = 1 \end{equation} | (A.29) |
we have
\begin{equation} p^\alpha({\pmb \xi}) = p^a({\bf x}) \, k^{\ \alpha}_{a}({\bf x}, {\pmb \xi}) \end{equation} | (A.30) |
that represents the unit vector parallel-transported from the point {\bf x} to the point {\pmb \xi} . In fact, since
\begin{equation} p^\alpha({\pmb \xi}) = p^a({\bf x}) e_{(i)}^{\ \alpha}({\pmb \xi}) e^{(i)}_{a}({\bf x}) \end{equation} | (A.31) |
we have in the triad basis
\begin{equation} p^{(i)}({\pmb \xi}) = p^\alpha(\xi^\alpha) \, e^{(i)}_{\ \alpha}({\pmb \xi}) = p^a({\bf x}) \, e^{(i)}_{a}({\bf x}) = p^{(i)}({\bf x}) \end{equation} | (A.32) |
and thus the components of the vector p^\alpha({\pmb \xi}) in the common triad basis coincide with those of p^a({\bf x}) . From this result, and from the property
\begin{equation} v^a({\bf x}) \, p_a({\bf x}) = v^a({\bf x}) \, g_a^{\ \alpha}({\bf x}, {\pmb \xi}) \, p_\alpha({\pmb \xi}) = v^\alpha({\pmb \xi}) \, p_\alpha({\pmb \xi}) \end{equation} | (A.33) |
it follows that any vector v^a({\bf x}) can be the parallel transported from {\bf x} to {\pmb \xi}' via the relation
\begin{equation} v^\alpha({\pmb \xi}) = v^a({\bf x}) \, g_a^{\ \alpha}({\bf x}, {\pmb \xi}) \end{equation} | (A.34) |
To conclude, an important bitensor is the so called Synge's world function [26], that is a measure of the geodetic distance between the points {\bf x} and {\pmb \xi} , defined as
\begin{equation} W({\bf x}, {\pmb \xi}) = \dfrac{1}{2}\int_{\pmb \xi}^{\bf x} ds^2 \end{equation} | (A.35) |
In Euclidean spaces, it can be explicited as as
\begin{equation} W({\bf x}, {\pmb \xi}) = \dfrac{1}{2}I_{(ij)}(x^{(i)}-\xi^{(i)})(x^{(j)}-\xi^{(j)}) \end{equation} | (A.36) |
x^{(i)} and \xi^{(i)} being the Cartesian coordinates of the two points. Its derivative at {\bf x} is
\begin{equation} W_a({\bf x}, {\pmb \xi}) = (x^{(i)}-\xi^{(i)}) \, I_{(ij)}\dfrac{\partial X^{(j)}}{\partial x^a} = ({\bf x}-{\pmb \xi})_a \end{equation} | (A.37) |
while the corresponding derivative at {\pmb \xi} reads
\begin{equation} W_\alpha({\bf x}, {\pmb \xi}) = (\xi^{(i)}-x^{(i)}) \, I_{(ij)}\dfrac{\partial X^{(j)}}{\partial \xi^\alpha} = ({\pmb \xi}-{\bf x})_\alpha = -\delta^a_\alpha({\bf x}-{\pmb \xi})_a \end{equation} | (A.38) |
Let ({\bf x}-{\pmb \xi})_\alpha = g^{\ a}_{\alpha}({\pmb \xi}, {\bf x})({\bf x}-{\pmb \xi})_a and ({\pmb \xi}-{\bf x})_a = g^{\ \alpha}_{a}({\bf x}, {\pmb \xi}) ({\pmb \xi}-{\bf x})_\alpha . From Eq (A.38) we have
\begin{equation} ({\bf x}-{\pmb \xi})_{\alpha_1}... ({\bf x}-{\pmb \xi})_{\alpha_n} = (-1)^n({\pmb \xi}-{\bf x})_{\alpha_1}... \\({\pmb \xi}-{\bf x})_{\alpha_n}; \qquad ({\pmb \xi}-{\bf x})_{a_1}... ({\pmb \xi}-{\bf x})_{a_n} = (-1)^n({\bf x}-{\pmb \xi})_{a_1}... ({\bf x}-{\pmb \xi})_{a_n} \end{equation} | (A.39) |
that is a useful relation in moment analysis.
In this Appendix we show that, owing to the property of the Stokes flow, the moments defined in Section 4 can be reduced to surface integrals of the stress tensor, in the case no-slip boundary conditions are assumed.
To this purpose, consider the following Stokes problems: Ⅰ) a flow with no-slip conditions at the boundaries of the fluid, {\bf v}({\bf x}) = {\bf w}({\bf x}) + {\bf u}({\bf x}) , where {\bf u}({\bf x}) is the ambient and {\bf w}({\bf x}) the disturbance flow. Thus, {\bf v}({\bf x}) is solution of the system
\begin{equation} \begin{cases} -\nabla \cdot {\pmb \sigma}({\bf x}) = \mu \Delta_{x} {\bf v}({\bf x})-\nabla p({\bf x}) = -{\bf f}({\bf x})\\ \nabla \cdot {\bf v}({\bf x}) = 0 \qquad {\bf x} \in \mathbb{R}^3\\ {\bf v}({\bf x}) = 0; \qquad {\bf x} \in \partial V_f\\ {\bf v}({\bf x}) = {\bf u}^\infty({\bf x}); \qquad {\bf x} \rightarrow \infty \end{cases} \end{equation} | (B.1) |
II) a purely n-th order velocity field {\bf u}_{(n)}({\bf x}) defined as in Section 5
\begin{eqnarray} u^{(n)}_a({\bf x}) = A_a^{\ {\bf a}_n}({\bf x} -{\pmb \xi}) _{{\bf a}_n} = A_\alpha^{\ {\pmb \alpha}_n} g_a^{\ \alpha}({\pmb \xi},{\bf x}) g_{{\pmb \alpha}_n}^{\ {\bf a}_n}({\pmb \xi}, {\bf x}) ({\bf x}-{\pmb \xi}) _{{\bf a}_n}\, , \\ \\ A^a_{\ {\bf a}_n} \, \nabla_a \left [ ({\bf x}-{\pmb \xi}')^{{\bf a}_n} \right ] = 0 \, , \quad {\bf x} \in \mathbb{R}^3, {\pmb \xi} \in V_{ext} \end{eqnarray} | (B.2) |
where A_a^{\ {\bf a}_n} are the components of a constant (n+1)-rank matrix, and the associated pressure p_{(n)}({\bf x}) and stress tensor {\pmb \pi}_{(n)}({\bf x}) are given by the Eqs (5.2) and (5.3).
Appling to the fields {\bf v}({\bf x}) an {\bf u}_{(n)}({\bf x}) the Lorentz reciprocal theorem [4]
\begin{equation} u_{a}^{(n)}({\bf x})\nabla_b \sigma ^{a b}({\bf x})- v_{a}({\bf x})\nabla _b \pi^{ab}_{(n)}({\bf x}) = \nabla_b(u^{(n)}_a({\bf x}) \sigma^{ab}({\bf x})- v_a({\bf x}) \pi^{ab}_{(n)}({\bf x})) \end{equation} | (B.3) |
Since these fields are defined in the whole space \mathbb{R}^3 , we can integrate Eq (B.3) over the domain of the obstacle V_{ext}
\begin{equation} \int_{V_{ext}} \left [ v_{a}^{(n)}({\bf x}) \, \nabla_b \sigma ^{a b}({\bf x})- v_{a}({\bf x}) \, \nabla _b \sigma^{ab}_{(n)}({\bf x}) \right ] \, \sqrt{g({\bf x})} d^3 x =\\ \int_{V_{ext}} \left [ \nabla_b(v^{(n)}_a({\bf x}) \, \sigma^{ab}({\bf x})- v_a({\bf x}) \, \sigma^{ab}_{(n)}({\bf x})) \right ] \, \sqrt{g({\bf x})} d^3 x \end{equation} | (B.4) |
Since \nabla_b\sigma^{ab}_{(n)}({\bf x}) = 0 , using Eqs (B.1), (B.2) and the definition of moments M_{\ {\pmb \alpha}_n}^\alpha({\pmb \xi}) in (4.4), the term at l.h.s in Eq (B.4) provides
\begin{equation} \int_{V_{ext}} f^a({\bf x}) v_a^{(n)}({\bf x}) \sqrt{g({\bf x})}d^3x = A_a^{\ {\bf a}_n} \\\int_{V_{ext}} f^a({\bf x})({\bf x}-{\pmb \xi})_{{\bf a}_n} \sqrt{g({\bf x})} d^3x = A_\alpha^{\ {\pmb \alpha}_n}M_{\ {\pmb \alpha}_n}^\alpha({\pmb \xi}) \end{equation} | (B.5) |
whereas, using Gauss divergence theorem the r.h.s, further considering that {\bf v}({\bf x}) = 0 and u^{(n)}_a({\bf x}) = A_a^{\ {\bf a}_n}({\bf x}-{\pmb \xi}) _{{\bf a}_n} on the boundaries \partial V_{ext} , one obtains
\begin{equation} \int_{V_{ext}} \left [ \nabla_b(v^{(n)}_a({\bf x}) \, \sigma^{ab}({\bf x})- v_a({\bf x}) \, \sigma^{ab}_{(n)}({\bf x})) \right ] \,\\ \sqrt{g({\bf x})} d^3 x = A_a^{\ {\bf a}_n} \int_{\partial V_{ext}} ({\bf x} -{\pmb \xi}) _{{\bf a}_n} \, \sigma^{ab}({\bf x}) n_b({\bf x})dS({\bf x}) \end{equation} | (B.6) |
Gathering Eqs (B.5) and (B.6), we have
\begin{equation} M_{\ {\pmb \alpha}_n}^\alpha({\pmb \xi}) = -\int_{\partial V_f} g^{\alpha}_{\ a}({\pmb \xi}, {\bf x}) g_{{\pmb \alpha}_n}^{\ {\bf a}_n}({\pmb \xi}, {\bf x}) ({\bf x} - {\pmb \xi})_{{\bf a}_n} \sigma^{ab}({\bf x}) n_b ({\bf x})dS \end{equation} | (B.7) |
which provides the expression of the moments in terms of surface integrals of the stress tensor.
[1] | W. Chen and M. Wang, Qualitative analysis of predator-prey models with Beddington-DeAngelis functional response and diffusion, Math. Comput. Model., 42 (2005), 31–44. |
[2] | M. Fan and Y. Kuang, Dynamics of a nonautonomous predator-prey system with the Beddington- DeAngelis functional response, J. Math. Anal. Appl., 295 (2004), 15–39. |
[3] | S. B. Hsu, T. W. Hwang and Y. Kuang, Global analysis of Michaelis-Menten type ratio-dependent predator-prey system, J. Math. Biol., 42 (2003), 489–506. |
[4] | N. Min and M. Wang, Hopf bifurcation and steady-state bifurcation for a Leslie-Gower preypredator model with strong Allee effect in prey, Discrete Cont. Dyn-A., 39 (2018), 1071–1099. |
[5] | W. Ni and M. Wang, Dynamics and patterns of a diffusive Leslie-Gower prey-predator model with strong Allee effect in prey, J. Differ. Equations, 261 (2016), 4244–4274. |
[6] | X. Yan and C. Zhang, Stability and turing instability in a diffusive predator-prey system with Beddington-DeAngelis functional response, Nonlinear Analysis RWA, 20 (2014), 1–13. |
[7] | F. Yi, J. Wei and J. Shi, Bifurcation and spatiotemporal patterns in a homogeneous diffusive predator-prey system, J. Differ. Equations, 246 (2009), 1944–1977. |
[8] | J. Ohlberger, O. Langangen, N. C. Stenseth, L. A. Vollestad, Community-level consequences of cannibalism, Am. Nat., 180 (2012), 791–801. |
[9] | V. H. W. Rudolf, M. Kamo and M. Boots, Cannibals in space: the coevolution of cannibalism and dispersal in spatially structured populations, Am. Nat., 175 (2010), 513–524. |
[10] | A. Basheer, E. Quansah, S. Bhowmick, et al., Prey cannibalism alters the dynamics of Holling- Tanner-type predator-prey models Nonlinear Dyn., 85 (2016), 2549–2567. |
[11] | L. V. Shevtsova and G. A. Zhdanova, V.A. Movchan, A.B. Primak, Experimental interrelationship between Dreissena and planktic invertribates, Hydrobiol. J., 11 (1986), 36–39. |
[12] | S. Chakraborty and J. Chattopadhyay, Effect of cannibalism on a predator-prey system with nutritional value: a model based study, Dyn. Syst., 26 (2011), 13–22. |
[13] | M. A. Elgar and B. J. Crespi, Cannibalism: Ecology and Evolution Among Diverse Taxa, Oxford University Press, New York, 1992. |
[14] | G. A. Polis, The evolution and dynamics of intraspecific predation, Annu. Rev. Ecol. Syst., 12 (1981), 225-251. |
[15] | S. Fasani and S. Rinaldi, Remarks on cannibalism and pattern formation in spatially extended prey-predator systems, Nonlinear Dyn., 67 (2012), 2543–2548. |
[16] | M. Genkai-Kato and N. Yamamura, Profitability of prey determines the response of population abundances to enrichment, Proc. R. Soc. Lond. B., 267 (2000), 2397–2401. |
[17] | M. Genkai-Kato, Nutritional value of algae: A critical control on the stability of Daphnia-algal systems, J. Plankton Res., 26 (2004), 711–717. |
[18] | C. Kohlmeier and W. Ebenhöh, The stabilizing role of cannibalism in a predator-prey system, Bull. Math. Biol., 57 (1995), 401–411. |
[19] | G. Q. Sun, G. Zhang, Z. Jin, et al., Predator cannibalism can give rise to regular spatial pattern in a predator-prey system, Nonlinear Dyn., 58 (2009), 75–84. |
[20] | J. B. Collings, Bifurcation and stability analysis of a temperature-dependent mite predator-prey interaction model incorporating a prey refuge, Bull. Math. Biol., 57 (1995), 63–76. |
[21] | C. B. Huffaker and C. E. Kennett, Experimental studies on predation: predation and cyclamen-mite populations on strawberries in California, Hilgardia, 26 (1956). |
[22] | L. Chen, F. Chen and L. Chen, Qualitative analysis of a predator-prey model with Holling type II functional response incorporating a constant prey refuge, Nonlinear Analysis RWA, 11 (2010), 246–252. |
[23] | A. P. Maiti, B. Dubey and J. Tushar, A delayed prey-predator model with Crowley-Martin-type functional response including prey refuge, Math. Method. Appl. Sci., 40 (2017), 5792–5809. |
[24] | J. P. Tripathi, S. Abbas and M. Thakur, Dynamical analysis of a prey-predator model with Beddington-DeAngelis type function response incorporating a prey refuge, Nonlinear Dyn., 80 (2015), 177–196. |
[25] | F. Wei and Q. Fu, Hopf bifurcation and stability for predator-prey systems with Beddington- DeAngelis type functional response and stage structure for prey incorporating refuge, Appl. Math. Model., 40 (2016), 126–134. |
[26] | Y. Du and J. Shi, A diffusive predator-prey model with a protection zone, J. Differ. Equations, 229 (2006), 63–91. |
[27] | X. Chang and J. Wei, Stability and Hopf bifurcation in a diffusive predator-prey system incorporating a prey refuge, Math. Biosci. Eng., 10 (2013), 979–996. |
[28] | R. Cui, J. Shi and B. Wu, Strong Allee effect in a diffusive predator-prey system with a protection zone, J. Differ. Equations, 256 (2014), 108–129. |
[29] | Y. Du, R. Peng and M. Wang, Effect of a protection zone in the diffusive Leslie predator-prey model, J. Differ. Equations, 246 (2009), 3932–3956. |
[30] | L. N. Guin and P. K. Mandal, Effect of prey refuge on spatiotemporal dynamics of the reactiondi ffusion system, Comput. Math. Appl. 68 (2014), 1325–1340. |
[31] | X. Guan, W. Wang and Y. Cai, Spatiotemporal dynamics of a Leslie-Gower predator-prey model incorporating a prey refuge, Nonlinear Analysis RWA, 12 (2011), 2385–2395. |
[32] | X. He and S. Zheng, Protection zone in a diffusive predator-prey model with Beddington- DeAngelis functional response, J. Math. Biol., 75 (2017), 239–257. |
[33] | Y. Cai, Z. Gui, X. Zhang, et al., W. Wang, Bifurcations and pattern formation in a predator-prey model, Int. J. Bifurcat. Chaos, 28 (2018), 1850140. |
[34] | K. D. Prasad and B. Prasad, Biological pest control using cannibalistic predators and with provision of additional food: a theoretical study, Theor. Ecol., 11 (2018), 191–211. |
[35] | A. M. Turing, The chemical basis of morphogenesis. Philos. Trans. R. Soc. Lond. Ser. B, 237 (1952), 37–72. |
[36] | C. G. Jäger, S. Diehl and M. Emans, Physical determinants of phytoplankton production, algal stoichiometry, and vertical nutrient fluxes, Am. Nat., 175 (2010), 91–104. |
[37] | R. Han and B. Dai, Spatiotemporal pattern formation and selection induced by nonlinear crossdi ffusion in a toxic-phytoplankton-zooplankton model with Allee effect, Nonlinear Analysis RWA, 45 (2019), 822–853. |
[38] | M. Scheffer and S. Rinaldi, Minimal models of top-down control of phytoplankton, Freshwater Biol., 45 (2000), 265–283. |
[39] | W. Wang, X. Gao, Y. Cai, et al., Turing patterns in a diffusive epidemic model with saturated infection force, J. Franklin I., 355 (2018), 7226–7245. |
[40] | Y. Lou and W. Ni, Diffusion vs cross-diffusion: an elliptic approach, J. Differ. Equations, 154 (1999), 157–190. |
[41] | C. Lin,W. Ni and I. Takagi, Large amplitude stationary solutions to a chemotaxis system, J. Differ. Equations, 72 (1988), 1–27. |
[42] | L. Nirenberg, Topics in nonlinear functional analysis, Courant Institute of Mathematical Science, New York, 1973. |
[43] | R. Peng, J. Shi and M. Wang, On stationary patterns of a reaction-diffusion model with autocatalysis and saturation law, Nonlinearity, 21 (2008), 1471–1488. |
1. | Mohammad Ferdows, Jahangir Alam, Ghulam Murtaza, Efstratios E. Tzirtzilakis, Shuyu Sun, Biomagnetic Flow with CoFe2O4 Magnetic Particles through an Unsteady Stretching/Shrinking Cylinder, 2022, 8, 2312-7481, 27, 10.3390/magnetochemistry8030027 | |
2. | Faris Alzahrani, O. Anwar Bég, M. Ferdows, Numerical solutions for magneto-convective boundary layer slip flow from a nonlinear stretching sheet with wall transpiration and thermal radiation effects, 2024, 85, 1040-7782, 1922, 10.1080/10407782.2022.2163945 |